Sunteți pe pagina 1din 115

Iowa State University

Digital Repository @ Iowa State University


Graduate Teses and Dissertations Graduate College
2013
Incorporating stochastic analysis in wind turbine
design: data-driven random temporal-spatial
parameterization and uncertainty quantication
Qiang Guo
Iowa State University, qguo@iastate.edu
Follow this and additional works at: htp://lib.dr.iastate.edu/etd
Part of the Mechanical Engineering Commons
Tis Dissertation is brought to you for free and open access by the Graduate College at Digital Repository @ Iowa State University. It has been accepted
for inclusion in Graduate Teses and Dissertations by an authorized administrator of Digital Repository @ Iowa State University. For more
information, please contact hinefuku@iastate.edu.
Recommended Citation
Guo, Qiang, "Incorporating stochastic analysis in wind turbine design: data-driven random temporal-spatial parameterization and
uncertainty quantication" (2013). Graduate Teses and Dissertations. Paper 13206.
Incorporating stochastic analysis in wind turbine design: data-driven random
temporal-spatial parameterization and uncertainty quantication
by
Qiang Guo
A thesis submitted to the graduate faculty
in partial fulllment of the requirements for the degree of
DOCTOR OF PHILOSOPHY
Major: Mechanical Engineering
Program of Study Committee:
Baskar Ganapathysubramanian, Co-major Professor
Jonathan Wickert, Co-major Professor
Eugene Takle
Frank Peters
Pranav Shrotriya
Iowa State University
Ames, Iowa
2013
Copyright c _ Qiang Guo, 2013. All rights reserved.
ii
DEDICATION
I would like to dedicate this thesis to my beautiful and wonderful wife Yaqin Liu who always
help me to become better and stronger. Without her unmitigated support in every possible
way I would not have been able to accomplish this work. I would also like to thank my friends
and family for their loving guidance and support during my study and research life.
iii
TABLE OF CONTENTS
LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi
LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x
ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
CHAPTER 1. OVERVIEW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Current Status of Wind Turbine Design . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 Status of wind modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Status of wind turbine modeling . . . . . . . . . . . . . . . . . . . . . . 4
1.1.3 Status of stochastic analysis on wind turbine . . . . . . . . . . . . . . . 7
1.2 Possible Improvements and Challenges . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 My Contribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
CHAPTER 2. TEMPORAL SPATIAL DECOMPOSITION: AN TURBU-
LENCE SIMULATION METHOD . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1 Mathematical Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.1 Stage 1: a low-dimensional representation via bi-orthogonal decomposition 14
2.1.2 Stage 2: Karhunen-Lo`eve expansion of spatial stochastic modes . . . . . 19
2.1.3 Stage 3: density estimation . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.1.4 Algorithm & implementation . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2 Discussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
CHAPTER 3. NUMERICAL EXAMPLES OF TSD . . . . . . . . . . . . . . . 26
3.1 Numerical Example 1: CWEX-11 . . . . . . . . . . . . . . . . . . . . . . . . . . 26
iv
3.1.1 Experiment setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.1.2 BD results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.1.3 KLE results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.1.4 The low-complexity wind model . . . . . . . . . . . . . . . . . . . . . . 37
3.1.5 Statistical comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2 Numerical Example 2: CWEX-10 . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2.1 Experiment setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2.2 Result: CO
2
uptake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.3 Numerical Example 3: Full-eld Stochastic Wind Simulation . . . . . . . . . . 49
3.3.1 Experiment setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.3.2 Result: statistical comparison . . . . . . . . . . . . . . . . . . . . . . . . 51
3.4 Discussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
CHAPTER 4. WIND TURBINE SIMULATION BASED ON STOCHAS-
TIC WIND . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.1 NREL 5MW Wind Turbine Model and FAST . . . . . . . . . . . . . . . . . . . 57
4.2 Uncertainty Quantication Using Adaptive Sparse Grid Collocation Method . . 60
4.3 Computational Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.4.1 Time responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.4.2 Convergence analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.5 Discussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
CHAPTER 5. WIND TURBINE SIMULATION BASED ON DETERMIN-
ISTIC WIND AND STOCHASTIC-ISOGEOMETRIC APPROACH . . . 73
5.1 Representing Randomness in Wind Turbine Blade Material . . . . . . . . . . . 73
5.2 Wind Turbine Blade Model Based on Isogeometric Analysis . . . . . . . . . . . 76
5.3 Computational Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.4.1 Stress analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
v
5.4.2 Failure analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.5 Discussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
CHAPTER 6. SUMMARY AND DISCUSSION . . . . . . . . . . . . . . . . . 92
BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
vi
LIST OF TABLES
2.1 Choices for Bi-orthogonal Decomposition. . . . . . . . . . . . . . . . . 18
2.2 Steps of computational framework. . . . . . . . . . . . . . . . . . . . . 24
3.1 KLE results of the rst three spatial modes . . . . . . . . . . . . . . . 36
3.2 Relationship between the number of random spatial modes (94% accu-
racy) and the number of random inputs. . . . . . . . . . . . . . . . . . 40
4.1 Wind turbine rotor geometry denition. [37] . . . . . . . . . . . . . . . 59
4.2 Normalized MSEs of dierent computational levels with respect to level
6, m = 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.3 Number of random samples at convergent level for the case of m = 1, 2, 3, 4. 69
5.1 Properties of berglass that is used. . . . . . . . . . . . . . . . . . . . . 76
5.2 Normalized MSE of dierent computational levels with respect to level 6. 84
vii
LIST OF FIGURES
1.1 Possible improvements in stochastic wind turbine analysis. . . . . . . . 10
3.1 Setup of experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2 10 minute period, averaged across all 144 time intervals and 28 realizations 29
3.3 Covariance function C(t, t

) . . . . . . . . . . . . . . . . . . . . . . . . 30
3.4 Spectrum of C(t, t

) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.5 First three eigenvectors of temporal covariance function . . . . . . . . 32
3.6 Stochastic spatial modes of C(t, t

) . . . . . . . . . . . . . . . . . . . . 33
3.7 Covariance function C
(2)
(t, t

) . . . . . . . . . . . . . . . . . . . . . . . 34
3.8 Spectrum of C
(2)
(t, t

) . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.9 First three eigenvectors of C
(2)
(t, t

) . . . . . . . . . . . . . . . . . . . 34
3.10 Stochastic spatial modes of C(t, t

) . . . . . . . . . . . . . . . . . . . . 35
3.11 The low-complexity model: Process of constructing synthetic wind snap-
shots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.12 Realizations of the stochastic ow . . . . . . . . . . . . . . . . . . . . . 39
3.13 PSDs and coherences of the synthetic turbulence at 10 m using one,
three, and ten modes compared with original ow . . . . . . . . . . . . 40
3.14 Comparing PSDs and coherences of the synthetic turbulence using dif-
ferent choices of snapshot . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.15 Comparison of covariance functions of original (a) and synthetic (b) ows 41
3.16 Kinetic Energy of Spatial Modes. . . . . . . . . . . . . . . . . . . . . . 42
3.17 Correlation of random energy of spatial modes. . . . . . . . . . . . . . 43
viii
3.18 Mean components of the CO
2
concentrations at upwind and downwind
towers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.19 Temporal covariance of CO
2
surface concentration . . . . . . . . . . . . 47
3.20 Eigenvalues (left) and cumulative fraction (right) of the rst 20 modes. 48
3.21 The rst temporal mode of CO
2
surface concentration . . . . . . . . . 49
3.22 15 15 grid on the 150 m 150 m rotor plane (yz plane) of the NREL
5MW wind turbine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.23 Covariance function of the TurbSim simulated turbulence. . . . . . . . 52
3.24 Spectrum of covariance matrix, (a) eigenvalues, (b) cumulative fraction
of energy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.25 Mean of the rst four stochastic spatial modes. . . . . . . . . . . . . . 53
3.26 Reconstructed time series of the along-wind turbulence at the rotor
center compared with the original TurbSim simulated ow. . . . . . . . 54
3.27 PSDs of the synthetic turbulence at rotor center using 2, 5, and 10 modes 55
3.28 Coherence spectra (calculated using 2, 5, and 10 BD modes) between
hub center p
0
and three lateral positions p
1
, p
2
, and p
3
(see Fig. 3.22). 55
4.1 Schematic of how FAST operates with AeroDyn and their input/output
les. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2 Airfoil cross-sections used in the design of the wind turbine rotor blades. [37] 59
4.3 Illustration of quantities from Table 4.1. [37] . . . . . . . . . . . . . . . 59
4.4 Schematic of incorporating TSD into full-eld wind simulation. . . . . 63
4.5 Time response of power generation by using 2, 5, and 10 modes and
TurbSim wind (top to bottom). . . . . . . . . . . . . . . . . . . . . . . 65
4.6 Time response of apwise bending moment at the half span of blade by
using 2, 5, and 10 modes and TurbSim wind (top to bottom). . . . . . 65
4.7 Time response of edgewise bending moment at the half span of blade
by using 2, 5, and 10 modes and TurbSim wind (top to bottom). . . . 66
ix
4.8 Time response of fore-aft bending moment at tower base by using 2, 5,
and 10 modes and TurbSim wind (top to bottom). . . . . . . . . . . . 66
4.9 PSD of apwise bending moment at half span of the blade. . . . . . . . 67
4.10 PSD of edgewise bending moment at half span of the blade. . . . . . . 67
4.11 PDFs of out-of-plane bending moment at blade root result from dierent
SGC computation levels. . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.12 PDFs of edgewise bending moment at half-span of blade, m = 2. . . . 70
4.13 PDFs of apwise bending moment at half-span of blade, m = 2. . . . . 70
4.14 PDFs of out-of-plane bending moment at blade root based on the tur-
bulent inows that are constructed by 1, 2, 3, and 4 BD spatial modes. 71
5.1 Wind turbine blade cross section. . . . . . . . . . . . . . . . . . . . . . 74
5.2 Surface meshes of the NREL 5MW wind turbine blade. . [9] . . . . . . 79
5.3 (a) Volumetric NURBS mesh of the computational domain and (b) A
planar cut to illustrate mesh grading toward the rotor blade. [9] . . . . 79
5.4 Simulation setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.5 PDFs at a critical location of dierent SGC computation levels. . . . . 83
5.6 Sampling points allocated by adaptive SGC algorithm (a) and non-
adaptive SGC algorithm (b) at computation level 6. . . . . . . . . . . . 84
5.7 Blade deformation under 11.4 m/s wind load and 12.1 rpm rotating speed. 85
5.8 Mean stress contour the 13
th
ply. . . . . . . . . . . . . . . . . . . . . . 86
5.9 Contour of the standard deviation of stress on the 13
th
ply. . . . . . . 86
5.10 PDF of the stress at the point that has the maximum standard deviation
of apwise stress. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5.11 Critical region that could have stress higher than 68MPa with large
probability. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5.12 Contour of failure index on the 14
th
ply. . . . . . . . . . . . . . . . . . 90
5.13 Probability density function of failure index at the point that has max-
imum mean failure index on the blade. . . . . . . . . . . . . . . . . . . 91
x
ACKNOWLEDGEMENTS
I would like to take this opportunity to express my appreciations to those who helped me
with various aspects of conducting research and the writing of this thesis.
First, I would like to gratefully thank Dr. Baskar Ganapathysubramanian for his guidance,
patience and support throughout this research and the writing of this thesis. His insights
and words of encouragement have often inspired me and renewed my hopes for completing
my graduate education. His passionate and strict attitude towards research will always be a
lifetime reference for me.
I would also like to sincerely show my gratitude to my co-major professor Dr. Jonathan
Wickert for oering me the opportunity to study in the US, leading me to my research area,
showing me the way to do great research, and supporting me throughout my graduate study.
I would like to thank my committee members for their eorts and contributions to this work.
I really appreciate Dr. Eugene Takle who provided me experimental data and the opportunity
to learn from a dierent area of research; Dr. Frank Peters who broadened my vision with his
great experiences in wind industry; and Dr. Pranav Shrotriya who supported me all the way
through my internship career where I gained precious experiences in wind industry. Without
your guidances and help, I would have been lost somewhere in the middle of my research career.
Finally, I would like to thank my lab mates for helping me with every subtle detail in my
research and creating this excellent research group where everyone is like my family.
xi
ABSTRACT
Wind turbines reliability is aected by stochastic factors in the turbulent inow and wind
turbine structures. On one hand, the variability in wind dynamics and the inherent stochastic
structures result in random loads on wind turbine rotor and tower. On the other hand, the
inherent structural uncertainties caused by imperfect control of manufacturing process intro-
duce unpredictable failures and decreases wind generators availability. Therefore, to improve
reliability, it is important to incorporate the variability in wind dynamics, and the inherent
stochastic structures in analyzing and designing the next generation wind-turbines.
In order to perform stochastic analysis on wind turbine, there are several improvements
need to be made. Current stochastic wind turbine analyses are mostly based on incomplete
turbulence input models. These models either failed to account for temporal variation of the
stochastic wind eld or unable to preserve spatial coherence which is a very important property
that describes turbulence structure. On the subject of modeling wind turbine, most commonly
used wind turbine design code is based on stead, lumped component blade models which lack
the ability to describe the complex 3D uid-structure interaction (FSI); which is essential to
provide precise blade stress distribution and deformation details. Finally, when it comes to
analyzing simulation results, most of existing work are done by analyzing the time response of
wind turbine, without looking at the stochastic nature of performance of wind turbines, and
its relationship between stochastic sources in turbulent inow and turbine structure.
In this work, we rst develop a data driven temporal and spatial decomposition (TSD),
which is capable of modeling any given large wind data set, to construct a low-dimensional yet
realistic stochastic wind ow model. Results of several numerical examples on the TSD model
show good consistency between given measured data and simulated synthetic turbulence. After
that, a stochastic simulation based on TSD simulated full-eld turbulence and a simplied wind
turbine model is performed. The result of this analysis shows the adequacy of using TSD as
xii
turbulence simulation tool as well as the random nature of wind turbines performance. Finally,
a stochastic analysis on a full scale 3D rich-structural wind turbine model with stochastic
composite material properties is performed. With a given steady wind load, the model gives
the deformation and the stress distribution of the blades. Critical regions that are most likely
to have stress larger than design strength of the material were identied. Failure analysis is
then performed based Tsai-Wu failure criterion.
1
CHAPTER 1. OVERVIEW
Wind energy is a clean and sustainable energy resource that is a promising alternative to
fossil fuel based energy. Great progresses have been made in wind power technology in the past
three decades. Wind turbine capacity has developed from less than 100 kW in the 1980s to a
current capacity of 5 MW. Wind turbine rotor diameter has also increased from 15 meters to
120 meters. Through improvements in the design of various subsystems of the wind turbine, the
energy production cost has been signicantly reduced. With these developments, large-scale
application of wind power has become a reality [25]. As a pioneer in wind power technology,
Denmark produced 22% of its electricity by means of wind power in 2010. In Asia, China added
16.5 GW of wind capacity in 2010 to reach 42.2 GW at the end of 2010 , which makes the
country the largest wind power market in the world [24]. U.S. Department of Energy (DOE)
has developed a scenario for supplying 20% of the nations electricity by means of wind by
2030 [66].
However, there are signicant challenges that have to be resolved in order to achieve these
goals. In fact, during the twelve months from July 2011 to June 2012, only 3.2% of all electricity
in the US was from wind [67]. One of the reasons for this sluggish use of wind is the current high
cost of wind energy, which can be markedly reduced by improving the eciency and reliability
of wind turbines [41]. As part of the focus on improving wind turbine eciency, both rotor
diameter and turbine hub hight have been increased signicantly. To minimize the vibration
caused by gravity, thinner and lighter blades are needed for these large scale wind turbines,
which will in turn aect their reliability. As a result, the need to decrease the weight of blades
while ensuring reliability becomes an important issue.
The reliability of a wind turbine is aected by its inherent material and operating un-
certainty/randomness. Examples of these random factors include turbulence inow, random
2
property of composite material from which blades are made, and random location and severity
of structural defects in blades. These uncertainties in wind turbine system inevitably aect the
reliability of wind turbines. Therefore, the reliability analysis of wind turbines requires a rig-
orous incorporation of the eect of these randomness. This research focuses on developing and
implementing new methods that incorporating stochastic analysis in wind turbine simulation.
1.1 Current Status of Wind Turbine Design
A wind turbine design work based on computational simulations could be done in three
steps: representing the wind inow as input of the wind turbine system, modeling the wind
turbine, and solving time responses of variables of interest as the simulation result. Reviews
on these three subjects are given in the following sections.
1.1.1 Status of wind modeling
The wind load on the turbine is a stochastic process whose direction and speed depends on
location, and time. In other words, ... at any instant they are distributed irregularly in space,
at any point in space they uctuate chaotically in time, and at given point and a given time
they vary randomly from realization to realization. [74] This randomness in wind conditions
leads directly to the uctuations in rotating speed of the rotor (for variable-speed wind turbine).
Furthermore, randomness of wind direction causes random yaw motion of nacelle, which can
further aect wind turbines productivity. Turbulence introduces unsteady loads on the blades.
All of these eects excite structural vibrations, introduce components failures, and further
reduce the lifespan of wind turbine. Thus, stochastic wind loads have a critical impact on wind
turbine performance.
Most of the past wind turbine design work was based on simplied wind load models
that assume steady wind speed, constant wind speed gradient prole and constant turbulence
intensity [51]. This kind of simplied wind model does not provide any insight into the eect of
randomness in the wind load. They have, however, been used with some success in the context
of deterministic wind turbine analysis [7, 8]. Subsequently, wind models that are capable of
describing wind stochasticity have been developed. Gaussian and Weibull distributions have
3
been used in approximating histograms of hourly mean wind speeds. These methods are based
on parameter estimation techniques such as maximum likelihood method (MLM) [12]. While
useful is representing randomness, these techniques lack the ability to describe temporal and
spatial correlations of wind, or location-dependent characterization capabilities. This results
in the development of methods that incorporate spatial variations, spatial correlation, and
temporal correlation.
There have been several seminal works that deal with preserving either spatial correlation
or temporal correlation. One such example is Veers/Sandia method [68]. The method rst
uses empirical coherence function and wind speed power spectral densities (PSDs) to describe
the wind eld. Choleski decomposition is then used to decompose the cross spectrum matrix.
Velocity time-series at dierent locations are nally calculated with a inverse Fourier transform
process. Another example is Manns method [44], which starts with a spectral tensor that
is achieved by taking incompressibility into account and solving the Navier-Stokes equations.
The wind eld is described in wave vector space and then transformed into spatial domain.
Recently, proper orthogonal decomposition (POD) was used in [55] to simulate wind turbine
inow. In this method, stochastic wind ow is viewed as wind eld snapshots at dierent
instances. A spatial covariance matrix is decomposed to get multiple characteristic modes
that are further used in constructing stochastic wind eld. By performing one POD analysis
at each time step, the wind ow is reproduced. Although the above three methods are easy
to implement and ecient enough to get random wind ows, they do not account for both
temporal correlation and spatial correlation of the wind ow at the same time. In software
package TurbSim [34], a turbulence simulation code developed by National Renewable Energy
Laboratory (NREL), Veers method is used. However, due to its inadequacy in describing
temporal coherence, additional information about coherent structure has to be included in the
model so that the turbulent structure in atmosphere could be more accurately represented [39].
most of existing stochastic wind models use assumptions of the certain stochastic proper-
ties of wind such as spatial coherence and temporal covariance, which may not be valid for
all applications. For analysis that focus on general wind conditions, such methods are ecient
and accurate enough. As the techniques of wind measurement developing, site specic high
4
frequency data will become available. In this case, data driven stochastic wind models are
desired for next generation wind turbine design. In fact, data driven wind models do exist.
Messac et al. [47] developed a wind model that is customizable based on local geographical
and climatic conditions. They used a nonparametric model to characterize the uncertainties
in annual multivariate distribution of wind speed and orientation so that the model could be
used in various locations. This model has been used to evaluate wind farm performance [79].
Although the model contains sucient information for the purpose of annual wind farm per-
formance evaluation, the diurnal variation of wind speed is not accounted for. Another method
that accounts for diurnal variability in wind speed was developed by Negra et al., where 10
minute average wind speed data over 7 years was used [49]. This method generates synthetic
wind speeds that are excellent statistical matches with the original data, but there is no easy
way to extend the method to describe spatial variations. It is worth noting that none of above
methods is able to model both spatial and temporal covariances at the same time.
Based on above discussion, there is a need of a more realistic parameterization of the wind
that encodes location-, topography-, diurnal-, seasonal and stochastic aects, and at the same
time be able to reproduce spatial and temporal covariances in the eld measured wind data.
However, such a comprehensive data driven parameterization is useful in practice only if it is
relatively simple (low-dimensional), which is the bottleneck of data driven wind models.
1.1.2 Status of wind turbine modeling
There is a trade o between improving eciency and increasing reliability of wind turbine.
On one hand, in order to improve wind turbine eciency, both rotor diameter and turbine hub
height have to be increased signicantly. On the other hand, to minimize the vibration caused
by gravity, thinner and lighter blades are needed for these large scale wind turbines, which
will in turn aect their reliabilities. A pressing challenge is to lighten the weight of blades and
simultaneously achieve high system reliability.
The reliability of a wind turbine is aected by material and operating uncertainty/randomness.
The stochasticity in the composite wind turbine blades depends on the type of composite ma-
terial used, the placement of composite ber pieces, the design of structural reinforcement
5
members such as shear web and spar caps, and structure parameters such as ply thickness.
Blades are made of composite laminate that has several plies. Each ply consists of a layer of
unidirectional bers, whose thickness and material properties can vary from ply to ply. The
randomness in composite material is introduced by the manufacturing process where a large
number of reinforcing elements need to be embedded into matrix. The resulting microscopic
heterogeneity impacts random ber positions, occurrence of defects in ber, etc. and causes
random behavior of the berglass ply under external loading [40]. Randomness in wind turbine
structure and material causes non-uniform rotor deformation that further results in unbalanced
load on the rotor. The random vibration introduced by this unbalanced load will decrease wind
turbines eciency and reliability. Therefore, in order to improve reliability of large scale o-
shore wind turbines, a crucial rst step is to have a better understanding of how they are
aected by manufacturing uncertainties. To this end, a comprehensive wind turbine model
that are able to incorporate structural and material uncertainties is required.
Present wind turbine modeling techniques can be basically classied into three types: nite
element methods (FEM), multi-body-system (MBS), and the modal approach [43]. A good
review of existing wind turbine design codes in these three classications are given in [2, 50]. In
the modal approach, modal analysis of beam-structured nite element wind turbine model is
performed rst. The rst few natural modal shapes are then superimposed to get the displace-
ment of turbine blade under certain give load. A good example is BLADED [21] which has
been an industrial standard tool for several years. However, it can not describe the torsional
blade deformations. This problem is solved in the MBS method, where the wind turbine system
is divided into a nite number of rigid/exural bodies connected with elastic joints. On each
body 2-dimensional loads (lift, drag and aerodynamic moment) are applied at the aerodynamic
center of the airfoil. A few ordinary/partial dierential equations are used to describe the be-
havior of the system. One examples of such design codes is FAST[36]. MBS method is fast and
reasonably accurate when general deections of all rigid bodies are sucient enough for simple
analysis. However, by simplifying the 3D geometry to 2D airfoil cross sections, the method fails
to capture the FSI between wind and rotor blades surface, which is essential for calculating the
detailed deection of the entire blade surface. In addition, for the purpose of structure design,
6
MBS method is insucient since it can not describe the eect of local structural changes on
system outputs. To this end, FEM is introduced in wind turbine design. In FEM, the detailed
blade geometry is discretized into nite elements, and the governing equations of the whole sys-
tem is discretized into a set of dierential equations. In contrast to the other two approaches,
FEM calculates accurate results but at high computational cost.
1
Current wind turbines are upwind turbines whose rotors facing towards the wind
2
. In this
case, blades may bend too much to hit the tower. To prevent such failure, the clearance between
the rotor blades and turbine tower (tip-tower clearance) should not be less than the minimum of
30% for the rotor turning and 5% for load cases with the rotor standing still, in relation to the
clearance in the unloaded state [18]. Three techniques are used to achieve this goal. One could
move the rotor further from the tower or change the angle of the rotor so that the bottom
of the rotor is further from the tower. However, the increased rotor overhang will increase
load on the shaft which may further reduce the reliability of the gearbox. A better approach
could be such that the blade is bent in the opposite direction of the blade bending moment
when unloaded. This process is called pre-bent or pre-coning [22]. This makes simulating wind
turbines more dicult because a pre-bent 3D wind turbine blade model is needed to fully assess
the impact of the new design to the blade stress as well as its performance. Although FEM
is very computational costly, it seems that it is the only choice among above methods when
performing structure design of rotors with complex geometry.
Most of existing wind turbine design codes assume small deections so that the aerodynamic
loads can be applied to the undeformed structure. However, this assumption is less valid since
modern wind turbine blades are more and more exible as their size increases. Given stable wind
load, the deformations of the exible blade inuences the aerodynamic force on blade surface
and vice versa. Therefore, analysis of oshore wind turbine blade is actually an aeroelastic
problem. To calculate the aeroelastic response, following aspects must be taken into account:
aerodynamic forces, blade deformations and the coupling between them [71]. By running the
analysis (with small displacement assumption) in a loop where the aerodynamic forces and
1
High computational cost is regarded as the major limitation of FEM.
2
Rotor of downwind turbine would suer strong turbulent which is caused by the uid structure interaction
between air ow and the tower, and generate large noise[53].
7
blade shape are updated in every iteration, the coupling problem can be solved. The only
remaining issue is that the FEM result is approximation of the analytic solution, which means
a small deviation from the beginning of the calculation could propagate to a very large one
after several iterations. Therefore, a mesh update and renement process is necessary in order
to minimize the calculation error in each iteration. Using FEM to solve aeroelastic problem for
exible wind turbine becomes an very inecient and cumbersome approach.
As an improvement of FEM, isogeometric analysis [28] provides much more accurate and
ecient geometric modeling. It represents exact geometries at the coarsest level of discretization
and eliminates geometrical errors from the beginning. The isogeometric method has been shown
to get accurate simulation results for exible wind turbine. A realistic 3D rotor model should
have the ability to describe the complex 3D uid-structure interaction (FSI) which will provide
us precise blade deformation and stress distribution [7, 8]. In addition, to assess the aects of
randomness in turbine structure to its performance, the wind turbine model must be able to
incorporate randomness in its structure and material. The advantages of isogeometric analysis
make it the ideal method for stochastic wind turbine analysis.
1.1.3 Status of stochastic analysis on wind turbine
Many turbulence models that are capable of describing long-term (10 minute average wind
speed) and short-term (spatial or temporal correlations) wind stochasticity have been devel-
oped. They have been successfully utilized in many applications. For instance, [47] developed
a wind model that is customizable based on local geographical and climatic conditions. They
used a nonparametric model to characterize the uncertainties in annual multivariate distri-
bution of wind speed and orientation so that the model could be used in various locations.
This model has been used to evaluate wind farm performance [79]. Simulation of wind turbine
aeroelastic response to wind turbine inow turbulence is given in [1]. In [59], fatigue failure
analysis is done with random wind load and 3-D nite element wind turbine model.
It must be mentioned here that almost all wind turbine simulation works based on stochastic
wind inputs are done by generating time responses of certain quantities of interest of wind tur-
bines. These results only describe how wind turbines operate given specic wind inputs. They
8
provides no information about how the randomness in turbulence input aect the performance
of wind turbines. To solve this problem, stochastic analysis/uncertainty quantication(UQ)
has to be incorporated in wind turbine design.
Stochastic analysis provides strategies to evaluate the relationship between a stochastic
input and various quantities of interest. In general, stochastic analysis is done by modeling
the system of interest, incorporating uncertainties in system model or its inputs, and getting
probability of the system output. For instance, to quantify the eect of the uncertainty that is
introduced by the random wind conditions, stochastic analysis (like polynomial chaos [77, 78],
stochastic collocation [16], or Monte Carlo analysis [20]) on a wind turbine model can be
performed [42].
By including uncertainty quantication in the modeling process, system governing equa-
tions (mostly a set of dierential equations) possess stochastic property as well. Such kind of
equations are called stochastic dierential equations (SDEs). Many methods have been devel-
oped to solve SDEs. Generally, these methods can be classied as statistical and non-statistical
methods [16]. One of the most commonly used statistical methods is Monte Carlo sampling.
In this method, realizations of random inputs are generated based on their probability distri-
bution. For each realization the data is xed and the problem is deterministic. A very large
number of samples are needed to get accurate result. Monte Carlo method quickly becomes
impractical when dealing with higher order complicated systems.
Non-statistical methods approximate the uncertainties in the system or its input rst, which
will be further used in system modeling. One example of this kind of methods is spectral
stochastic nite element method (SSFEM) [20]. In this technique, the random eld is ex-
panded about its mean with a set of complete orthogonal polynomials whose coecients are
realizations of a set of random variables. Generalized polynomial chaos (gPC) [77] is another
well developed non-statistical methods. With gPC, stochastic solutions are expressed as orthog-
onal polynomials of the input random parameters. Followed by Galerkin projection, the set of
stochastic dierential equations become deterministic decoupled equations which can be solved
by using common numerical methods. Many orthogonal bases could be used to construct the
polynomials. Hermite polynomial expansion is of most commonly used. In addition, a whole
9
family of polynomials named Wiener-Askey polynomial chaos can be used according to dierent
distribution functions of the random inputs [77]. In this way, the optimal convergence can be
achieved. Similar to Monte Carlo method, non-statistical methods also convert the stochastic
problem to a set of deterministic equations. However, the resulting equations generated by
non-statistical methods are often coupled, which makes using this method to solve complicated
system very dicult.
Most recently, stochastic collocation approach is used to solve SDEs [4]. This method is
based on deterministic sampling method like Monte Carlo method. Instead of nding the
solution of a huge amount of sampling points in the random eld and then form a discrete
solution of the random input, one can use less sampling points, get the deterministic solutions,
and then use these solutions to form an approximate collocation solution. How to choose as
less as sampling points to reduce the computational cost and, at the same time, meet the
prescribed accuracy is the main concern of stochastic collocation method. One way is to use
tensor products of one-dimensional Gaussian quadrature points as sampling points [4]. As
the dimension of random elds increase, the number of sampling points grows exponentially
fast. Sparse grid collocation (SGC) technique is introduced in order to reduce the number of
sampling points in higher random dimensions [76]. Another issue of stochastic collocation is how
to choose basis functions to interpolate the collocation solutions. Lagrange polynomials and,
more general, gPC polynomials are used. Adaptive sparse grid collocation method is further
developed by introducing the concept of adaptability in to the algorithm [16]. SPG method is a
very ecient method that has its advantages in solving complicate, large stochastic dimensional
problems.
1.2 Possible Improvements and Challenges
Fig. 1.1 shows dierent subroutines in wind turbine simulation work and possible improve-
ments with respect to incorporating stochastic eects in the analysis. As the source of energy,
inow wind plays an important role in wind turbine simulation. Based on discussion in previous
section, a data driven wind model that is able to reproduce temporal and spatial covariances
of the source wind is in great need. However such model has to be low-complexity and easy to
10
use.
The second type of improvements could be done in wind turbine modeling. A full-scale, 3D
comprehensive wind turbine model is required. Such model would provide detailed deections
and accurate surface stress distributions, which are very important for successful wind turbine
blade design. A developing work from [7, 8] is very promising. The diculties of such model
lies in incorporating stochastic parameters, such as random defects, material properties, and
stochastic parameters, into the complicated wind turbine structure.
Improvements could also be done in interpreting simulation results. To improve wind tur-
bine reliability, a crucial step is to have a better understanding of how they are aected by
stochastic sources of the system such as wind input and random defects. This can be accom-
plished using stochastic analysis. Since wind turbine simulation is a very complicated and time
consuming process, even getting one set of simulation results would require huge amount of
computational eort. It becomes very dicult to perform stochastic analysis in wind turbine
design, where a large number of simulations need to be done.
FSI Solver
Structure
Solver
Stochastic Wind
Generator
Stochastic
Structure
Force u,v,w
Aeroelastic Solver
Stochastic
Results
Material Properties
Random Defects
Stochastic Parameters

Data Driven Model
Temporal Covariance
Spatial Coherence

Surface Stress
Detailed Deflection
Probability Results

Figure 1.1 Possible improvements in stochastic wind turbine analysis.
1.3 My Contribution
The contributions of this work are listed as follows.
1. Stochastic wind modeling:
(a) formulated a mathematical framework that is able to accurately represent wind ow
using a low-complexity yet realistic stochastic model (see Chapter 2);
11
(b) implemented a computational framework based on the mathematical framework that
extracts statistical information from any given turbulence data and constructs an
easy-to-use model (see Chapter 2);
(c) applied the computational framework to three numerical examples (see Chapter 3).
2. Stochastic analysis of wind turbine:
(a) used the stochastic wind ow that was developed in this work performed wind turbine
simulation on a simplied wind turbine model (see Chapter 4);
(b) performed stochastic analysis on the simplied wind turbine model using adaptive
sparse grid collocation method (see Chapter 4);
(c) Incorporated stochastic material properties into a full-scale, 3D, complex geometry,
wind turbine rotor model (see Chapter 5);
(d) performed stochastic analysis, stress analysis, and failure analysis on the 3D wind
turbine model (see Chapter 5).
12
CHAPTER 2. TEMPORAL SPATIAL DECOMPOSITION: AN
TURBULENCE SIMULATION METHOD
The motivation for this chapter is the fact that no single existing wind simulation method
can provide a data-driven model that seamlessly accounts for spatial variations, diurnal vari-
ations, and temporal correlations. To this end, an ecient method that leverages the huge
amount of meteorological data that is readily available will be very useful for various aspects
of wind turbine analysis. In this chapter, we develop a data-driven space-time parameteriza-
tion of any given large data set of wind to construct a low-dimensional yet realistic stochastic
wind ow model. The framework is based on a two-stage model reduction: Bi-orthogonal De-
composition (BD) followed by Karhunen-Lo`eve expansion (KLE). The resulting time-resolved
stochastic model encodes most of the statistical properties in the given wind ow. Moreover,
the temporal modes encode the variation of wind speed in the mean sense and resolve temporal
correlation while the spatial modes provide deeper insight into spatial of the wind eld - which
is a key aspect in current wind turbine sizing, design and classication. In addition, several
interesting ramications of this low dimensional model are discussed. These include informa-
tion about the energy cascade, which is computed as correlations between random energy of
dierent spatial modes.
In the following part, section 2.1 focuses on constructing a low-complexity model of the
random wind ow. Bi-orthogonal Decomposition (BD) [70, 69] and Karhunen-Lo`eve expansion
(KLE) [20] are used in constructing the low-dimensional wind model. The algorithmic details
of the computational framework is give in section 2.1.4. Finally, section 2.2 concludes this
chapter.
13
2.1 Mathematical Framework
Denote by x = (x
1
, x
2
, x
3
) = (x, y, z) the three-dimensional coordinates with x, y, and z rep-
resents along-wind, transverse, and vertical axis respectively, Wind velocity v = (v
1
, v
2
, v
3
) =
(u, v, w) consists of three wind speed components along the three spatial axes. A stochastic
wind can be dened as v(x, t, ), where =
i
, i = 1, . . . , n is a random vector associated
with the random eld. It is worth noting that the random eld described above is actually a
function v : X T R where X R
d
(d = 1, 2, 3) denotes the spatial domain, T R
denotes the temporal domain and is the sample space of a set of random variables that is
related to the random eld.
The analysis is clearer when the mean component v(x) of the velocity data is removed so
that only the uctuation components u(x, t, ) are left, i.e.
u(x, t, ) = v(x, t, ) v(x), . (2.1)
The mean component is dened as
v(x) =
1
[T[
_
T
v(x, t, )dt. (2.2)
where denotes the average in the stochastic domain, [T[ is the span of the temporal domain,
v is the ensemble average.
The goal of this work is to construct a simple stochastic model that encodes
temporal and spatial covariance, and preserve all the statistical information, such
as spatial coherence and wind speed power spectral density (PSD), present in the
collected eld measured data. We look for a model that has the form
u =

K
ijk
X
i
(x) T
j
(t)
k
(2.3)
where K
ijk
are coecients, the deterministic functions X
i
track spatial correlations, T
j
track
temporal correlations and the random variables
k
encode the inherent variability. The goal
becomes to nd the simplest possible representation that still encodes all the required informa-
tion.
14
The main idea of the next section is to formulate a mathematical strategy of representing
this meteorology data in terms of the smallest possible number of terms in Eqn. 2.3, by opti-
mally designing X, T, and . Noting that the data contains spatial, temporal and stochastic
variabilities, we solve this problem in three stages. In the rst stage, we decompose the data
into temporal (T(t)) and coupled spatial-stochastic ((x, )) parts through the concept of Bi-
orthogonal Decomposition; in the second part we decompose the spatial-stochastic part into
spatial (X(x)) and stochastic () components using the concept of the Karhunen-Lo`eve decom-
position; the third stage focuses on estimating probability density functions of TSD generated
random variables.
2.1.1 Stage 1: a low-dimensional representation via bi-orthogonal decomposition
In the rst stage, we are looking for a minimal representation of the data in the form
u(x, t, ) =

i,j
K
ij

i
(x, ) T
j
(t) (2.4)
where i and j are independent indices. Consequently, the expression on the right hand side has
an dramatically large number of terms. This is handled utilizing the Schmidt decomposition
theorem [57], which states that any representation of a tensor product space H = H
1
H
2
can be expressed as linear combination of tensor product of basis functions
i

i
, where

i
H
1
,
i
H
2
. As a result, the representation can be reduced to
u(x, t, )
M

i=1
K
i

i
(x, ) T
i
(t). (2.5)
where
i
(x, ) are stochastic spatial modes and T
i
(t) are temporal modes.
Our goal becomes searching for the best choices for
i
and T
i
such that the decomposition
uses the least number of terms M that will give us an accurate representation of u(x, t, ).
We pose this as an optimization problem. To do so, we dene the error in this representation
and design
i
and T
i
that minimize this error. The error, denoted by , is dened as the
low-complexity stochastic model subtracted from the true data
(x, t, ) = u(x, t, )
M

i=1
K
i

i
(x, ) T
i
(t). (2.6)
15
Approximation theory suggest that the best choice for the functions
i
and T
i
(we will also call
them modes) is when they are orthogonal to each other [29]. Thus, we set T
i
, i = 1, . . . , M to
be orthogonal to each other in the time domain and
i
, i = 1, . . . , M to be weakly orthogonal
in spatial domain. Mathematically, this is denoted in terms of the inner products:
T
i
, T
j

T
=
_
T
T
i
(t)T
j
(t)dt =
ij
(2.7)
and

i
,
j

X
=
_
X

i

j
dx =
ij
, (2.8)
where
i
denotes the expectation of the spatial-stochastic mode, i.e.

i
(x) =
_

i
(x, )W() d, (2.9)
and W() is the multivariate joint probability density of random variables in the set .
Note that the error itself is an random eld. We construct an associated scalar value with
this random eld to accomplish subsequent optimization. To this end, an error functional E is
dened as the norm of , i.e.
E =
_
T
,
X
dt. (2.10)
The error-functional is simply the inner product (i.e. an average) of the eld over space, time
and stochastic dimensions. Note that error functional depends on the choice of functions T
i
(t)
and
i
(x, ), i.e. E [T
1
, , T
M
,
1
, ,
M
].
We now search for temporal functions that minimizes this error functional. This is ac-
complished by applying the calculus of variations and solving the associated Euler-Lagrange
equations [15]. We provide full details of the derivation as follows.
Theorem 1. (Eulers equation) Let J[y] be a functional of the form
_
b
a
F(x, y, y

)dx, (2.11)
dened on the set of functions y(x) which have continuous rst derivatives in [a, b] and satisfy
the boundary conditions y(a) = A, y(b) = B. Then a necessary condition for J[y] to have an
extremum for a given function y(x) is that y(x) satisfy Eulers equation
F
y

d
dx
F
y
= 0 (2.12)
16
The goal of BD is to minimize the error functional
E [T
1
, , T
M
] =
_
T
,
X
dt. (2.13)
Substituting representation error (Eqn. 2.6) in above equation yields
E [T
1
, , T
M
] =
_
T
_
u(x, t, )
M

i=1
K
i

i
(x, )T
i
(t), u(x, t, )
M

j=1
K
i

j
(x, )T
j
(t)
_
X
dt
(2.14)
According to the denition in Eqn. 2.8, by taking the spatial inner product, the randomness
in Eqn. 2.14 is removed by the ensemble average operation, the error functional becomes
E [T
1
, , T
M
] =
_
T
_
X
_
u(x, t)
M

i=1
K
i

i
(x)T
i
(t)
_
_
_
u(x, t)
M

j=1
K
j

j
(x)T
j
(t)
_
_
dx dt
=
_
T
_
X
u
2
(x, t) 2u(x, t)
M

i=1
K
i

i
(x)T
i
(t) +
M

i=1
K
i

i
(x)T
i
(t)
M

j=1
K
j

j
(x)T
j
(t) dx dt
=
_
T
_
X
u
2
(x, t) 2u(x, t)
M

i=1
K
i

i
(x)T
i
(t) +
M

i=1
K
2
i

2
i
(x)T
2
i
(t) dx dt
=
_
T
_
_
X
u
2
(x, t)dx 2
M

i=1
K
i
T
i
(t)
_
X
u(x, t)
i
(x)dx +
M

i=1
K
2
i
T
2
i
(t)
_
dt
=
_
T
F(t, T
1
, , T
M
, T

1
, , T

M
)dt
where
F(t, T
1
, , T
M
, T

1
, , T

M
) =
_
X
u
2
(x, t)dx 2
M

i=1
K
i
T
i
(t)
_
X
u(x, t)
i
(x)dx +
M

i=1
K
2
i
T
2
i
(t).
(2.15)
In above derivation, the orthogonality of basis functions
i
and T
i
is applied. According to
Theorem 1, the error functional E [T
1
, , T
M
] has extremum when T
i
satisfy Eulers equations
F
T
i

d
dx
F
T

i
= 0. (2.16)
That is
_
X
u(x, t)
i
(x)dx K
i
T
i
(t) = 0, (2.17)
17
where
T
i
(t) =
1
K
i
u(x, t, ),
i
(x, )
X
. (2.18)
Applying temporal inner product , T
i

T
to Bi-orthogonal Decomposition (Eqn. 2.5) and
considering the orthogonality of the temporal basis functions yields

i
(x, ) =
1
K
i
u(x, t, ), T
i
(t)
T
. (2.19)
Above two equations dene a coupled relationship between T
i
and
i
. Now that we have two
unknowns and two equations, they can be solved by substituting Eqn. 2.19 into Eqn. 2.18,
which results in eigenvalue problem for temporal modes
K
2
i
T
i
(t) =
_
T
C(t, t

)T
i
(t

)dt

, (2.20)
where C(t, t

) is called temporal covariance that can be obtained by taking the inner product
in spatial domain, i.e
C(t, t

) = u(x, t, ), u(x, t

, )
X
. (2.21)
Setting
i
= K
2
i
, Eqn. 2.20 can be transformed to

i
T
i
(t) =
_
T
C(t, t

)T
i
(t

)dt

, (2.22)
where
i
and T
i
(t) are eigenvalues and eigenfunctions of the covariance function C(t, t

). The
optimal choice of temporal modes T
i
and
i
can be obtained by solving the eigenvalue problem.
Once
i
and T
i
are solved for, the spatial-stochastic functions are calculated using

i
(x, ) =
1

i
u(x, t, ), T
i
(t)
T
. (2.23)
It should be mentioned that the rst M (usually M 36) eigenvalues and eigenfunctions
usually represent the data exceedingly well [16]. Thus, the rst stage of the decomposition of
the wind data results in representation involving temporal functions T
i
and spatial-stochastic
functions (x, ). Eqn. 2.5 now becomes
u(x, t, ) =
M

i=1
a
i
(x, ) T
i
(t), (2.24)
18
where
a
i
(x, ) = K
i

i
(x, ) =

i

i
(x, ). (2.25)
are spatial-stochastic modes.
Remark 1. We chose to decompose the data into spatial-stochastic and temporal parts in the
rst stage of the decomposition. This decomposition is one of three possible decomposition
choices. These choices are enumerated in Table 2.1 where X, T, and denote the spatial,
temporal, and stochastic domains respectively. Such decompositions have been explored in other
works. For instance, Venturi et al. [70, 69] investigated Type 1 decomposition. The decomposi-
tion suggested by Type 2 can be achieved by using generalized polynomial chaos [77, 78]. Type
3 decomposition is used in [46] to model uncertain cylinder wake. It can be shown that Type 1
and Type 3 decompositions will result in identical result.
Table 2.1 Choices for Bi-orthogonal Decomposition.
Type Space 1 Space 2
1 X T
2 X T
3 T X
Remark 2. The choice of the inner products (temporal Eqn. 2.7 and spatial-stochastic Eqn. 2.8)
aect the properties of the decomposition. In particular, there are several dierent ways in which
we can dene an inner product over the spatial-stochastic modes (i.e. dierent ways to average
over space and stochastic dimensions). These include the following possibilities:

i
,
j

0
=
_
X

i

j
dx,

i
,
j

1
=
_
X

i

j
dx,

i
,
j

2
=
_
X

i

j

i

j
dx,
The rst inner-product (denoted by
0
) is a spatial integral of the product of expected values,
while the second inner-product (denoted by
1
) is the expectation of the spatial integral.
It can be shown that by taking inner products,
h
, of type h = 0, 1, 2, we obtain optimal
representations with respect to mean, second-order moment, and standard deviation of the data,
19
respectively. We have chosen to focus on representation that is optimal in the mean sense [70].
2.1.2 Stage 2: Karhunen-Lo`eve expansion of spatial stochastic modes
In this stage, we decompose the spatial-stochastic functions a
i
(x, ) into a spatial part and
a set of uncorrelated random variables. First, the mean of spatial-stochastic functions a(x) are
removed so that only the uctuation component is left for analysis.
(x, ) = a(x, ) a(x). (2.26)
Following the rational of Bi-orthogonal Decomposition (Eqn. 2.5), our goal is to decompose
(x, ) into a minimal set of linear combination of deterministic spatial functions and uncor-
related random variables, that is
(x, )
N

i=1
C
i

i
X
i
(x). (2.27)
where C
i
are coecients of the expansion,
i
is a set of uncorrelated random variables, X
i
are deterministic spatial functions. We pose this decomposition problem as an optimization
problem, where the optimization problem is to minimize the error. The representation error is
dened as
= (x, )
N

i=1
C
i

i
X
i
(x) (2.28)
This is a standard formulation of the Karhunen-Lo`eve expansion. The goal of KLE is to nd
the optimal choice for functions X
i
such that the representation error is minimized with a
nite number (N) of expansion terms. We briey describe the mathematical framework of
KLE below [20]: The representation error is converted into a cost-functional for optimization
by simply considering the mean-square error (i.e. the inner-product)
E
2
=
_
X

2
dx. (2.29)
Minimization of the error-functional results in an eigenvalue problem, whose eigenfunctions are
the desired spatial functions X
i
_
X
R(x
1
, x
2
) X
i
(x
2
) dx
1
= C
2
i
X
i
(x
2
), (2.30)
20
where R(x
1
, x
2
) is the covariance kernel constructed from the spatial-stochastic functions
(x, ). We denote C
2
i
=
i
so that
i
and X
i
(x) are the eigenvalues and the eigenvectors of
the covariance kernel. Eqn. 2.27 becomes
(x, )
N

i=1
_

i
X
i
(x). (2.31)
In order to solve the eigen-problem (Eqn. 2.30), the covariance kernel must be given or
calculated from measured data. The Wiener-Khinchin theorem allows computing the covariance
in a very ecient way.
Given realizations of the spatial-stochastic mode, i.e. (x), the covariance, R, can be
computed as:
R = F
1
(F() F()

). (2.32)
where F()

is the complex conjugate of Fourier transform of , and the diagonal entries of


R contains the covariance.
To solve Eqn. 2.30 numerically, the equation should be discretized rst. To this end, eigen-
function can be approximated by linear combination of N basis functions [20]
X
k
(x) =
N

i=1
d
(k)
i
h
i
(x). (2.33)
Substitute above equation to the eigen-equation and set the error to be orthogonal to each
basis function yields
N

i=1
d
(k)
i
__
X
__
X
R(x
1
, x
2
)h
i
(x
2
)dx
2
_
h
j
(x
1
)dx
1

n
_
X
h
i
(x)h
j
(x)dx
_
= 0. (2.34)
Above equation can be written in matrices form
AD = BD. (2.35)
A
ij
=
_
X
_
X
R(x
1
, x
2
)h
i
(x
1
)h
j
(x
2
)dx
1
dx
2
. (2.36)
B
ij
=
_
X
h
i
(x)h
j
(x)dx =
_
X
H
T
(x)H(x)dx. (2.37)
D
ij
= d
(j)
i
. (2.38)

ij
=
ij

i
. (2.39)
21
where H(x) = (h
1
(x), h
2
(x), , h
N
(x)). Matrix A can be rewritten as
A =
_
X
_
X
H
T
(x
1
)R(x
1
, x
2
)H(x
2
)dx
1
dx
2
=
_
X
_
X
H
T
(x
1
)H(x
1
)RH
T
(x
2
)H(x
2
)dx
1
dx
2
=
_
X
H
T
(x
1
)H(x
1
)dx
1
R
_
X
H
T
(x
2
)H(x
2
)dx
2
= BRB.
and
R(x
k
, x
l
) =
N

i=1
N

j=1
h
i
(x
k
)R
ij
h
i
(x
l
)
= h
k
(x
k
)R
kl
h
l
(x
l
)
= R
kl
.
2.1.3 Stage 3: density estimation
The nal step of the formulation is to identify the probability distributions of the uncorre-
lated random variables
i
. Note that we have same number of realizations of
i
as the number
of random samples of the stochastic wind. Each realization is computed by inverting Eqn. 2.31
for
i
:

i
=
1

i
_
X
X
i
(x)dx. (2.40)
Density estimation is a process of estimating the probability density function (PDF) of an
undened stochastic process based on observed data. Many approaches to density estimation
have been developed, which can roughly be divided into two categories, namely parametric
estimation and non-parametric estimation. Maximum likelihood estimation (MLE)[11] and
Bayesian estimation[17] are two examples of parametric estimation. In MLE, the parameters
of the targeting distribution are assumed as xed but unknown. Then the parameters are
found with certain optimization method such that the resulting PDF best describes the data
observation. Bayesian estimation assumes the parameters are random variables with some
known priori distributions. The distribution of the undened process can be calculated using
theorem of total probability and Bayes rule. To perform these methods, whether the targeting
22
distribution or the prior density of parameter need to be known, which is usually not given or
easily constructed. In this case, non-parametric density estimation can be used.
The simplest form of non-parametric density estimation is histogram which has drawbacks,
such as discontinuity, dependence on the starting position, and the dependence of the number
of samples. To avoid these shortcomings, more advanced approaches have been developed. The
general expression for non-parametric density estimation is
p(x)

=
k
NV
, (2.41)
where V is the volume surrounding x, N is the total number of samples, k is the number of
examples inside V . Choosing a xed value of k and determine the corresponding volume V from
the data yields the k Nearest Neighbor (kNN) approach[14]. On the other hand, if choosing a
xed value of V and determine k from the data, Kernel Density Estimation (KDE) is resulted.
A good introduction of KDE could be found in[58].
With respect to addressing stochasticity in wind, all wind simulation methods can be divided
into two categories, i.e. parametric and nonparametric methods. All methods discussed in the
introduction section use parametric methods. In most cases, the probability density function
of the velocity uctuations in homogeneous turbulence is assumed to be Gaussian or sub-
Gaussian [31]. Although this assumption is often valid for most of the physical phenomena
(including wind at a at terrain) in nature, it is not necessarily true in the case of complicated
terrain. On the other hand, although nonparametric methods are robust and easy to implement,
its accuracy strongly depends on number of available data samples.
In this research, a nonparametric method, named kernel density estimation (KDE), is used
to describe the stochasticity of the turbulence. It is worth noting that the applicability of the
framework developed in this context should not depend on the choice of density estimation
techniques. The choice of technique to construct the probability distribution of
i
given a
nite number of observations of
i
is crucial. We utilize Kernel Density Estimation (KDE)
methods [58] to construct the probability distributions of
i
in a non-parametric way.
In KDE, the PDF of a random variable is estimated as
p() =
1
Nh
N

i=1
K
_

i
h
_
, (2.42)
23
where K() is the kernel which is a symmetric function that integrates to one, and h > 0 is
a smoothing parameter called the bandwidth. Common choices for K() are the multivariate
Gaussian density function. The choice of the bandwidth is critical since a small values of
h result in the estimated density with many wiggles, while large values of h result in very
smooth estimations that do not represent the local distributions. Several approaches have been
developed to chose proper h values. A good review on bandwidth selection can be found in [64].
We advocate using a simple formula based on Silvermans rule [60]. The optimal choice for h
according to Silvermans rule is given by
h =
_
4
3N
_1
5
1.06 n

1
5
, (2.43)
where is the standard deviation of the samples.
Based on above three techniques, we are now able to construct space-time decomposition of
the wind led snapshots and preserve the spatial and temporal correlations. In next chapter,
we look at an numerical example that illustrate the power of this framework.
2.1.4 Algorithm & implementation
The algorithmic details of the framework is outlined in Table 2.2. The matrix of wind
ow snapshots is rst constructed. The temporal covariance matrix is calculated from the
snapshot matrix and stored as a datale. Next, the eigenvalue problem (corresponding to the
Bi-orthogonal Decomposition) is solved to get temporal and spatial-stochastic modes. Fol-
lowing this, the spatial-stochastic modes are decomposed into spatial functions and uncorre-
lated random variables via the Karhunen-Lo`eve expansion. This is accomplished by solving an
eigenvalue problem. The eigenvectors computed are the spatial functions and they are used to
compute realizations of the uncorrelated random variables (via Eqn. 2.40). Finally, a Kernel
Density Estimator is used to construct the PDFs of random variables
i
j
. At this time, all the
necessary elements for constructing synthetic wind ow were known. Finally, synthetic wind re-
alizations are constructed by reversing above process. The complete framework is implemented
in C++. SLEPc [26], which is based on PETSc [5], was used in solving eigenvalue problems.
The complete framework can be downloaded from this URL [23].
24
Table 2.2 Steps of computational framework.
1. Data preparation
(a) Construct matrix of wind speed eld snapshots from given data
2. Bi-orthogonal Decomposition
(a) Calculate temporal covariance from snapshots matrix
(b) Solve eigenvalue problem, get eigenvalues and eigenfunctions
(temporal modes)
(c) Solve spatial stochastic modes
3. Karhunen-Lo`eve expansion
(a) Calculate covariance kernel of each spatial mode (M spatial modes
in total)
(b) Solve eigenvalue problem for each spatial mode, get basis spatial
functions
(c) Calculate observations of each random variable
i
j
based on
Eqn. 2.40
4. Kernel density estimation
(a) Estimate PDF for each random variable
i
j
based on observations
5. Construct synthetic wind ow
(a) Get random samples from random variables
i
j
(b) Construct stochastic spatial modes according to Eqn. 2.31
(c) Construct synthetic wind ow using Eqn. 2.5
25
2.2 Discussion and Conclusion
Incorporating the eects of randomness in wind is critical for a variety of application in-
volving wind energy. Continuous advances in data-sensing and meteorology has made possible
the availability of large data-sets of location-, topography-, diurnal-, and seasonal sensitive
meteorology data. While this data contains rich information, ease of use is bottlenecked by
the unwieldy data-sizes. A pressing challenge is to utilize this data to construct a location-
, topography-, diurnal-, and seasonal-, dependent low-complexity model that is easy to use
and store. We formulate a data-driven mathematical framework that is capable of represent-
ing the spatial- and temporal- correlations as well as the inherent randomness of wind into
a low-complexity parametrization. We leverage data-driven decomposition strategies like Bi-
orthogonal Decomposition and Karhunen-Lo`eve expansion for constructing the low-complexity
model. We provide a software package that can be used to construct the low-complexity model
to the community.
26
CHAPTER 3. NUMERICAL EXAMPLES OF TSD
Chapter 2 provides data-driven low-complexity turbulence simulation tool that is powerful
in representing location- and topography specic wind given comprehensive wind eld measure-
ment. In this chapter, we test the framework by providing three numerical examples. Analysis
on the examples include generating synthetic stochastic wind ow and comparing several wind
turbulence statistics between synthetic and original ows.
3.1 Numerical Example 1: CWEX-11
First, we illustrate the methodology based on a recent meteorological experiment named
CEWX-11. CWEX-11 is a collaborative experiment (Iowa State University (ISU) and the Uni-
versity of Colorado (UC), assisted by the National Center for Atmospheric Research (NCAR)).
CWEX-11 and its 2010 counterpart (CWEX-10) in the Crop Wind-energy EXperiment, ad-
dress observational evidence for the interaction between large wind farms, crop agriculture, and
surface-layer, boundary-layer, and mesoscale meteorology [52].
3.1.1 Experiment setup
In the experiment, a surface ux station was installed to the south of a wind turbine,
which makes the station measure upstream inow of the wind turbine due to the fact that
predominant summer winds in Iowa originate from south to slightly south-east. The surface
station was equipped with a CSAT3 sonic anemometer that was located at height 4.5 m and an
RMY propeller and vane anemometer that was located at height 10.0 m. The former measured
wind speed in 3 directions at 20 Hz whereas the latter gave wind speed amplitude and its
27
direction at 1 Hz. The experiment started on June 29 and lasted for 48 days.
1
A schematic of the experiment is shown in Fig. 3.1. Note that the gure is for illustrative
purpose and is not drawn to scale. In this experiment, only wind speed magnitude was analyzed,
but it is straightforward to perform analysis on any one of the three components of turbulence.
Stochastic
wind load
NCAR
surface flux
station
, v t x
z
x
y
( , ) x z x
Figure 3.1 Setup of experiment
Ideally, we would like to have multiple wind eld snapshots taken at regular intervals on a
vertical plane that is perpendicular to the rotor. However, due to the constrain of the experi-
ment, only two time-series were measured at heights 4.5 m and 10.0 m. Linear interpolation of
the two time-series at 18 spatial points on z direction were performed to simulate more vertical
measurements. The wind snapshots are constructed by using Taylors frozen turbulence hy-
pothesis [62]. Specically, all measurements in the resulting 20 time-series over certain interval
were treated as if they were taken at the same instant in the interval. This would provide us
along-wind (x direction) measurements. Finally, one 2D snapshot taken at certain instant is
constructed
2
. Similarly, snapshots of the wind eld at other instants were created to represent
one full day of measurements. We have data for 28 such days. The full meteorology data
1
Excluding the days that mostly have opposite wind direction and the days with sudden rain, only data for
28 days that had the desired weather condition and wind direction were used in the analysis.
2
Since no measurements were taken in transverse direction, only 2D wind eld is analyzed in this analysis.
28
curated into this form can be represented as following matrix
v
11
(x, z) v
12
(x, z) v
1m
(x, z)
v
21
(x, z) v
22
(x, z) v
2m
(x, z)
.
.
.
.
.
.
.
.
.
v
n1
(x, z) v
n2
(x, z) v
nm
(x, z)
In the matrix, each element v
ij
(x, z), i = 1, 2, , m; j = 1, 2, , n represents a snapshot
of wind eld. Each column contains snapshots that span an entire day. Each row represents
data for a particular interval measured over dierent days. Thus, each row of data can be
considered to be realizations of the stochastic wind eld at a particular interval.
It is noteworthy that the length of time interval that was used to obtain wind eld snapshots
can be arbitrarily chosen. Although, most guidelines for wind turbine design and siting suggest
considering wind variabilities over a period of 10 minutes for wind data analysis [18], in this
example, analysis based on dierent choices of time intervals were performed and compared.
In the results section, we will show that using 10 minutes time interval to construct wind eld
snapshots is a reasonable choice.
The meteorological data is provided in the common NetCDF (Network Common Data
Form) format. The original meteorological data contains a variety of information, including
wind speed and orientation, temperature, humidity, surface CO
2
ux. For this analysis we only
consider the wind speed data. The wind speed data is extracted from the NetCDF le using
MATLAB. TSD software package is then used to perform the decomposition and construct
synthetic wind ow.
3.1.2 BD results
As discussed previously, CWEX-11 meteorological data consists of wind speed amplitude
measured at 1 Hz over 28 days. Snapshots curated based on 10 minutes interval are used rst
and the reason of doing so is going to be explained in the following part. As a result, there are
144 snapshots in one day and we have 28 such days (samples). We will describe the stage wise
construction of the low-dimensional model and its accuracy in the next few subsections.
29
The mean component of the ow that is averaged across all 144 snapshots and 28 realizations
according to Eqn. 2.2 is shown in Fig. 3.2. This plot shows us the mean wind prole such as
vertical wind shear load. Note that the x-axis represents length of the snapshots that was
calculated based on 10 minute interval and average wind speed.
Figure 3.2 10 minute period, averaged across all 144 time intervals and 28 realizations
The temporal covariance are calculated based on Eqn. 2.21. By taking the inner products
across the spatial domain, we construct the covariance function C(t, t

) which is shown in
Fig. 3.3. Note that C has block structure, with regions of high covariance (marked by the solid
boxes) along the diagonal and regions of large negative covariance along the o-diagonal. This
structure of the covariance function follows the dynamics of stable, and unstable stratication
of the atmospheric boundary layer seen in the US central plains. We discuss this by dividing the
analysis into several distinct time periods (marked by the solid boxes). Following meteorological
practice, the data starts at Coordinated Universal Time (UTC, or Greenwich time) 00:00. The
rst period is UTC 00:00-06:00 (CST 18:00-00:00) which corresponds to the time between
sunset to midnight. In this period, insolation and, thus, heating is gradually cut o and the
temperature of atmosphere cools down (from the ground up) due to rapid cooling of the ground.
Because of the heavier density of the cooler air, the cooler air stays at the bottom (close to
the surface). This generates a stably stratied boundary layer that does not change during
the duration of the night. This results in the high covariance between adjacent time periods
marked in the lower left box. The second period is the region of reduced covariance between
UTC 06:00-13:00 which is basically the time from midnight to shortly after sunrise. During
30
summer, the nocturnal Great Plains low-level jet (LLJ) exists in Iowa. For a good overview
of LLJ, see [30]. The LLJ causes low level turbulence and enhanced mixing which reduces
the covariance between adjacent-time wind elds. The third time period is between UTC
13:00-00:00 which corresponds to sunrise and day time. Because of the sunrise, the ground is
rapidly heated. As a result, the warmer air near the ground becomes buoyant and rises rapidly
with its place being taken by compensating ow of colder air from higher up in the atmospheric
boundary layer. Eddies are thereby created, changing from smaller scale to larger scales, nally
becoming the circulatory motion that crosses the entire boundary layer so that the high speed
free stream ow is brought in the circulation. This phenomenon usually happens at noon,
which can also be found in Fig. 3.5 at UTC 18:00. In addition, the covariance function exhibits
periodic behavior that is caused by the periodical motion of eddies.
Daytime Sunset to midnight
Figure 3.3 Covariance function C(t, t

)
The temporal covariance function plays a very important role in this analysis since it pro-
vides almost all the needed information that describes the behavior of the wind ow. Compared
to the original meteorological data, the temporal covariance is much easier to store and trans-
mit, yet contains nearly the same amount of information as the original meteorological data.
31
The eigenvalues of the temporal covariance function are shown in Fig 3.4. Fig 3.4(a) compares
the relative magnitudes of eigenvalues; the magnitudes of the eigenvalues provide a notion of
how much energy about the data is stored in each spatial mode [27]. Notice that the rst
eigenvalue is much larger than the other subsequent eigenvalues, which means the rst mode
contains largest portion of the energy in the turbulence eld. In Fig 3.4(b) this is represented
as the cumulative fraction of energy contained in the rst k modes. This plot provides us with
a precise notion of how many terms are needed to incorporate [27], say, 90% of the information
available in the data into the low-complexity model. The rst ve eigenvalues cover about 90
% of the total energy of the turbulence eld, which ensures that a ve term decomposition
will have a 90% accuracy of representation. This illustrates the advantage of Bi-orthogonal
Decomposition. As a reduced order model, Bi-orthogonal Decomposition is able to represent
a random ow with much fewer random modes compared to the number of original snapshots.
In other words, by using BD, we reduced the terms in representing the given stochastic random
ow from 144 snapshots to a simple 5-variable parametrization.
(a) (b)
Figure 3.4 Spectrum of C(t, t

)
Fig. 3.5 shows the rst three eigenfunctions of C(t, t

). Note that the eigenfunctions are


the temporal functions T
i
. They track how the uctuation component of the random ow
varies during a day. Based on the plot, the largest mean (and standard deviation, which is
not reported here) of wind speed occurs at UTC 18:00 because of the reason stated previously.
32
From Fig. 3.4, we notice that the rst mode carries about 85% of the total energy in the
turbulence. Therefore, the trend of diurnal uctuation of the turbulence eld can be mostly
seen in the rst temporal mode, with the next two temporal modes look like white noise. Once
2

3

Figure 3.5 First three eigenvectors of temporal covariance function
eigenvalues and eigenvectors for temporal covariance function are solved, the spatial-stochastic
modes a
i
(x, ) can be constructed (Eqn. 2.23 and Eqn. 2.25). Fig. 3.6 shows the expectation
of the rst three spatial modes. It is clear that the rst spatial mode that carries the largest
part of turbulence energy describes the vertical shear pattern, while the second mode describes
the lateral shear pattern of the wind eld. Higher modes that account for more complicated
turbulence structures are insignicant since they contain limited turbulence energy.
Based on Remark 2, dierent choices of inner products result in optimal representation
with respect to dierent statistical properties of turbulence. For the sake of illustration, BD
results calculated by using inner product h = 2 are shown in Fig. 3.7, 3.8, 3.9, and 3.10. It
is worth noting that by choosing inner product
2
, we are solving the eigenvalue problem
with a dierent covariance kernel. The diagonal of Fig. 3.7 represents the (spatial) average
of the ow standard deviation [70]. It can be seen that the maximum standard deviation
occurs at UTC 18:00 which is the same time as maximum mean wind speed occurs (see in
Fig 3.3). In addition, Fig. 3.8 shows that in order to get 90% of representation accuracy with
respect to higher order statistics of turbulence, much more eigenvalues are needed. The rst
three eigenvectors (temporal modes) are shown in Fig. 3.9. Unlike Fig. 3.5 that corresponds
to the case of h = 0, where the second and third temporal modes are white noises, all three
temporal modes have apparent trends of wind speed uctuation. This is because the rst
33
Figure 3.6 Stochastic spatial modes of C(t, t

)
three eigenvalues only preserve 55% of total energy, which means none of them dominates the
turbulence. Similar result is shown in Fig. 3.10 all three spatial modes contribute to the vertical
sheer load.
Although analysis based on inner product type h = 2 shows many interesting results, when
it comes to the resulted synthetic wind ow, we found that it is very similar to the case of
h = 0. In fact, the two simulated ows only dier at the accuracies when representing the
random ow for dierent applications. Therefore, in the following context, only the results
using inner product h = 0 are shown.
3.1.3 KLE results
Using the technique introduced in section 2.1.2, the spatial-stochastic modes are decom-
posed into deterministic modes and random variables. KLE results of the rst three spatial
modes are shown in Table 3.1. Covariance matrices, eigenvalues, deterministic spatial modes,
and probably distributions of random variables are shown in the table from top to bottom. As
34
Figure 3.7 Covariance function C
(2)
(t, t

)
(a) (b)
Figure 3.8 Spectrum of C
(2)
(t, t

)
Figure 3.9 First three eigenvectors of C
(2)
(t, t

)
35
Figure 3.10 Stochastic spatial modes of C(t, t

)
discussed in Sec. 2.1.3, the distributions of
i
j
are estimated by using KDE.
From Fig. 3.6 we notice that as the mode order increases, the complexity of spatial mode
also grows. As a result, higher order spatial modes require more KLE terms to be accurately
represented. This statement can be veried by comparing results of the rst three spatial
modes. For instance, the smoothness of the covariance kernels, the complexity of deterministic
modes, and the required numbers of KLE terms to achieve 90% of representation accuracies all
increases as the mode number becomes higher.
It is worth to mention that the PDFs of
i
j
are not always Gaussian. Although Gaussian
assumption is commonly used in turbulence analysis, it is seen from Fig. 3.1 that Gaussian
assumption may not always be valid. Notice that each random sample corresponds to one day
wind history, the randomness of this long term stochastic process may be dierent from the
randomness of the more commonly used 10-minute average wind speed whose distribution is
always regarded as Gaussian or Weibull.
36
a
1
(
x
,

)
a
2
(
x
,

)
a
3
(
x
,

)
(
a
)

(
b
)

(
a
)

(
b
)

(
a
)

(
b
)

X
1

X
2

X
3

X
1

X
3

X
2

X
1

X
3

X
2

T
a
b
l
e
3
.
1
K
L
E
r
e
s
u
l
t
s
o
f
t
h
e

r
s
t
t
h
r
e
e
s
p
a
t
i
a
l
m
o
d
e
s
37
3.1.4 The low-complexity wind model
The above two stages result in the calculation of the temporal functions, T
i
(t) (from Bi-
orthogonal Decomposition), the spatial functions (X
i
j
(x)) (from KLE decomposition), and the
uncorrelated random variables,
i
j
(from the KDE tting).
3
Putting it all together gives us the
low-complexity model for the wind
u(x, t, ) =

K
ij
T
i
(t)X
i
j
(x)
i
j
(3.1)
Note that X
i
j
and T
i
are deterministic functions that encode spatial and temporal correlations of
the wind. Dierent realizations (or stochasticity) of the wind is included into the low-complexity
model via the uncorrelated random variables,
i
j
. Note that the probability distributions of
i
j
are constructed in a data-driven way from the meteorology data. As we will show in the results
section, only a few terms (M = 3) are required to reconstruct a synthetic wind snapshot
that contains all the temporal and spatial correlations exhibited by the original
data. Fig. 3.11 gives a graphical description of this process. Synthetic wind snapshots exactly
mimicking the meteorological data can be constructed by sampling from the distributions of
the random variables,
i
j
. Interestingly, only 800 KB of storage space is needed to store all the
necessary information of the reduced-order model, whereas more than 200 MB of storage space
is needed to store all the wind ow snapshots that are used in the analysis! This demonstrates
the advantage of the low-complexity model in terms of data size.
3.1.5 Statistical comparison
As discussed previously, to get 90% representation accuracy, ve terms in Bi-orthogonal
Decomposition and seven terms in Karhunen-Lo`eve Expansion are needed. However, for the
purpose of demonstration, 1-, 3-, and 10-term BD and three terms in KLE for each BD term
were used in the analysis. In order to quantify the accuracy, 28 realizations (same as the
number of samples of meteorological data set) of the synthetic wind ow are generated. Each
realization is a 24-hour wind data consisting of 144 ten-minute snapshots (see Fig. 3.11 for one
24 hour synthetic dataset).
3
The superscript in X
i
j
(x) and
i
j
represents the index of Bi-orthogonal Decomposition terms whereas the
subscript denotes the index of KLE terms.
38
10 minutes
24 hours
50 100 150 200 250 300 350 400 450 500 550 600
2
4
6
8
10
12
14
16
18
20
( ) a x
( ) v x
2
( ) X x
3
( ) X x
1
( , ) a x
Time/hours(UTC)
H
e
i
g
t
h
/
m

0 5 10 15 20
5
6
7
8
9
10
1
2
3
4
( , , ) v t x
1
( ) X x
=
=
Figure 3.11 The low-complexity model: Process of constructing synthetic wind snapshots
In Fig. 3.12, realizations of the stochastic ow at height 10.0 m using one, three, and ten
spatial modes are compared. Result shows that when smaller number of terms are used, the low
frequency characteristics of the turbulence can be represented accurately. However, in order to
describe high frequency behavior, higher modes in BD must be used.
One of the most important variables of turbulence is the power spectral density (PSD)
that describes how the energy in turbulence is distributed to dierent frequency spectrum.
To verify the similarity of the reconstructed wind ow and the meteorological data, their
PSD functions are compared in Fig. 3.13 (a). The gure shows that the synthetic wind ow
accurately reproduces energy in low frequency region. However, more terms need to be used
in the representation to preserve energy in high frequency region. It is worth to mention that
compare with POD method, even with only one BD mode, the synthetic wind mimics the true
data in great detail.
Coherence spectrum is another important variable that describes the similarity of turbulence
at two dierent spatial locations. Comparison of coherences between same set of time-series
used in PSD analysis is given in Fig. 3.13 (b). Note that using only one BD mode, the synthetic
wind can still preserve most of the spatial coherence information. This can be explained by
the advantages of STD. Since STD performs orthogonal decompositions of both temporal and
39
S
y
n
t
h
e
t
i
c

a
n
d

o
r
i
g
i
n
a
l

t
u
r
b
u
l
e
n
c
e
,

m
/
s

UTC time, hour
Original
10 modes
3 modes
1 mode
3

2.5
3

2.5
0 5 10 15 20
3

2.5
3

2.5
Figure 3.12 Realizations of the stochastic ow
spatial covariances, the temporal variations and spatial correlations are preserved in an optimal
way.
As previously mentioned, there are dierent choices of time intervals when constructing
wind ow snapshots. In above content, only analysis on 10 minute snapshot was performed.
The reason for choosing 10 minute snapshot can be demonstrated by comparing PSDs and
coherences of the synthetic turbulence using dierent choices of time interval averaging pro-
cesses. Fig. 3.14 shows results of such analysis. Since the accuracy of PSDs of synthetic wind
only depends on how many total terms in the decomposition, there is no apparent dierence
in PSDs of dierent choices of time intervals. On the other hand, comparison of coherences
tells us if performing analysis on snapshots that are constructed from less than 10 minute eld,
important information in turbulence coherence will be lost. Coherences correspond to 10, 15,
and 20 minute snapshots are very close, which means the additional turbulence data provided
by 15 and 20 minute snapshots, comparing to 10 minute snapshot, does not contain much more
coherence information. This result is consistent with GL guideline that states the coherence
structure in turbulence can be assumed to be unchanged in 10 minutes period of time [18].
According to Fig. 3.4, certain number of spatial modes are needed to achieve a desired
40
(a) (b)
Figure 3.13 PSDs and coherences of the synthetic turbulence at 10 m using one, three, and
ten modes compared with original ow
representation accuracy. On the other hand, the number of random inputs to the reduced order
model equals the number of spatial points on each snapshot. By using dierent choices of time
intervals, the resulting snapshots have dierent dimensions. Table 3.2 shows the relationship
between the number of random inputs and the number of needed spatial modes for certain level
of accuracy. Based on the table, when the number of random inputs increases by 20 times,
the needed random spatial modes to achieve 94% of accuracy only increases from 3 to 8. In
other words, the required number of spatial modes does not strongly depend on
the number of random inputs, which makes TSD to be practical for problems with
very large stochastic dimensions.
Table 3.2 Relationship between the number of random spatial modes (94% accuracy) and the
number of random inputs.
Number of random inputs 1200 6000 12000 18000 24000
Number of spatial modes 3 5 6 7 8
Finally, since the temporal covariance contains important temporal information, it is worth
to compare the temporal covariances of the source and synthetic winds. The temporal covari-
ance function of synthetic data

C(t, t

) is constructed by using the 28 simulated realizations


of the synthetic random ow. Comparing it with the temporal covariance of the original data
41
(a) (b)
Figure 3.14 Comparing PSDs and coherences of the synthetic turbulence using dierent
choices of snapshot
reveals that they have almost identical pattern. The temporal error (or the information loss)
(a) (b)
Figure 3.15 Comparison of covariance functions of original (a) and synthetic (b) ows
can be dened as the L
2
-norm of the dierence between the covariance functions:
=
| C

C |
2
| C |
2
=
_

n
i=1

n
j=1
[ C(t
i
, t
j
)

C(t
i
, t
j
) [
2

n
i=1

n
j=1
[ C(t
i
, t
j
) [
2
_
1/2
. (3.2)
where n = 144 is the number of snapshots in 24 hours. The temporal information-loss using
a 9-term expansion is 2.25%. Thus, a 9-term data-driven expansion reproduces the temporal
and spatial covariance of the original meteorological data to 97.75% accuracy.
42
The spatial and temporal functions used in the construction of the low-complexity model
provide signicant insight into the structure of the wind eld. For example, each spatial-
stochastic mode describes the wind eld at dierent grades of detail [27]. The (stochastic)
kinetic energy contained in each spatial-stochastic mode is dened as

i
=
1
2
_
X
[a
k
(x, )]
2
dx. (3.3)
This allows us to calculate the correlation between the kinetic energy content across dierent
spatial-stochastic modes
C
KE
(i, j) =

i

j

i

j

j
, (3.4)
where

i
is the standard deviation of the random energy of ith spatial stochastic mode. The
correlation provide information about how dierent spatial-stochastic modes exchange energy
(i.e. how kinetic energy cascades across length scales). A high correlation between two modes
means the energy transition between them is prominent. The mean value and standard devia-
tion of the kinetic energy of dierent spatial modes are shown in Fig. 3.16.
K
i
n
e
t
i
c

E
n
e
r
g
y

o
f

S
p
a
t
i
a
l

M
o
d
e
s
,

(
m
/
s
)
2

Figure 3.16 Kinetic Energy of Spatial Modes.
Correlation between the dierent spatial-stochastic modes is shown in Fig. 3.17. Interest-
ingly, the energy cascade occurs between consecutive and also some prominent non-consecutive
modes. For example, it can be seen that the energy transfer from both the rst and second
modes mainly transfer to the fourth mode, which explains the reason that the kinetic energy
of the fourth mode is larger than the third mode (see Fig. 3.16).
43
1.0

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
Figure 3.17 Correlation of random energy of spatial modes.
3.2 Numerical Example 2: CWEX-10
One of the challenges in wind farm commissioning is the social acceptance of onshore wind
turbines. W ustenhagen et al. dened three dimensions of social acceptance, namely socio-
political acceptance, community acceptance and market acceptance [73]. Specically, the com-
munity acceptance refers to the public opinion to wind energy project. Although people usually
have positive attitudes about wind energy in general, they tend to resist wind projects dur-
ing the actual on-site planning process with the so called NIMBY (Not In My Back Yard)
argument. In fact, the public acceptance of wind projects is inuenced by many factors. For
example, the nancial benet to the community, the impact of wind farm to public activities,
and the awareness of the local population to the benets and impacts of wind facilities [32].
The impacts of wind farm on environment, wildlife, and peoples ordinary life are among the
most concerns of the community. The impact of wind turbines on wildlife (birds and bats) has
been well studied in [6, 3]. Visual impact (disturbance) of wind turbine is investigated in [10].
A good review on wind turbine acoustic noise is given in [54].
On the other hand, despite the fact that most of the land-based wind resources are located
in the states that contain a lot of farmlands, the eect of wind farm on crops is still unclear.
It is important to investigate the upwind and downwind variation in air conditions to assess
44
the impact of wind turbines on crops. This necessitates a realistic parameterization of air ow
that encodes location-, topography-, diurnal-, seasonal and stochastic aects. However, such a
parameterization is useful in practice only if it is relatively simple (low-dimensional). Interest-
ingly, the data to construct such models are available at various resolutions from meteorological
measurements. Such meteorology data contain rich information about location- and topogra-
phy specic temperature, H
2
O, and CO
2
concentration data which are important factors in air
ow that would aect crops growth. In this section, we utilized the data-driven framework
developed previously to generate space-time parameterizations of the large scale meteorology
data measured at upwind and downwind meteorology towers. Comparing the two parameter-
izations gave us deeper understanding of how wind turbine may variate CO
2
concentration in
air ow.
3.2.1 Experiment setup
The data to construct the stochastic model of wind is based on meteorological data mea-
sured by the CWEX-10 that addresses observational evidence for the interaction between large
wind farms, crop agriculture, and surface-layer, boundary-layer, and mesoscale meteorology
(Rajewski et al. 2012) [52]. In the experiment, a surface concentration station (NLAE 1, where
NLAE is the abbreviation of National Laboratory for Agriculture and the Environment) was
installed 4.5 D (D denotes turbines rotor diameter) to the south of a wind turbine, which makes
the station measure upstream inow of the wind turbine due to the fact that predominant sum-
mer winds in Iowa originate from south to slightly south-east. Three downwind concentration
towers (NLAE 2, NLAE 3, and NLAE 4) were placed at 2.5 D, 17 D, and 35 D, respectively,
north of the wind turbine. Since the NLAE 3 and NLAE 4 did not measure H
2
O and CO
2
,
measurements of surface concentrations at height 6.45m from the NLAE 1 and NLAE 2 are
used in current analysis. The meteorological data consisted of high-speed (20 Hz) velocity,
temperature, CO
2
and moisture measurements upwind and downwind of a working 1.5 MW
turbine taken over two months.
Similar to the CWEX-11 example, the meteorology data is curated into a set of time-series
vectors. To exclude the eects of environmental factors to the wind eld and to make sure
45
NLAE 1 is upwind tower, days with rain, wrong wind direction, and invalid measurements are
taken out from the original data set. As a result, 15 days of measurements are left for analysis.
In this work, CO
2
surface concentrations data consisting of certain duration of measurements
are treated as one snapshot of the time-series. The full meteorology data curated into this form
can be represented as a matrix

11
(x)
12
(x)
1m
(x)

21
(x)
22
(x)
2m
(x)
.
.
.
.
.
.
.
.
.

n1
(x)
n2
(x)
nm
(x)
In the matrix, each element
ij
(x), i = 1, 2, , m; j = 1, 2, , n represents a snapshot
of surface concentration time-series. Each column contains snapshots that span an entire day.
Each row represents data for a particular time interval measured over dierent days. Thus, each
row of data can be considered to be realizations of the stochastic wind ow at a particular time
interval. This naturally motivates us to consider the wind ow as a one-dimensional random
eld with spatial and temporal variations, (x, t; ), where x, t, and represents the spatial
domain, temporal domain, and stochastic variability respectively.
The analysis is clearer when the mean component

(x) of the data is removed so that only
the uctuation components (x, t, ) are left, i.e.
(x, t, ) = (x, t, )

(x), (3.5)
where the mean component is dened as

(x) =
1
[T[
_
T
(x, t, )dt. (3.6)
and [T[ is the total time, i.e 24 hours, for one realization.
BD of is given in the following equation
(x, t, )

(x) +
M

i=1

i
(x, ) T
i
(t). (3.7)
where
i
(x, ), i = 1, . . . , M are stochastic spatial modes that are weakly orthogonal in the
time domain and T
i
(t) i = 1, . . . , M are temporal modes that are orthogonal in spatial domain,
and

i
are coecients.
46
In Eqn. 3.7,

(x) represents the mean component and the second term on the right hand
side represents the uctuation part of the quantity of interest in the ow. The variance of the
quantity of interest can be dened as
V ar() = E[(

)
2
]
= E
_
_
_
M

i=1

i
(x, ) T
i
(t)
_
2
_
_
=
M

i=1

i
(3.8)
where we utilized the orthogonality of temporal and spatial modes.
3.2.2 Result: CO
2
uptake
Currently, most guidelines for wind turbine design and siting suggest considering wind
variabilities over a period of 10 minutes for wind data analysis [18]. On the other hand, much
longer averaging period is accepted for agricultural and atmospheric analysis. This is because
changes in air ow conditions need about 20-30 minutes for the boundary layer to achieve a
quasi-equilibrium. The process may take much longer at night, which implies even 30-minute
average may not be long enough. Without loss of generality, 30-minute average is used in the
analysis.
Based on Eqn. 3.6, the mean component of CO
2
concentration is calculated by taking
the average on temporal and stochastic domain. Fig. 3.18 shows the comparison of mean
components of the upwind and downwind ows. The increment of the mean CO
2
concentration
is shown as 1.08% in the gure.
Fig. 3.19 shows the upwind temporal covariance with respect to 30-minute averaging pe-
riod. Following meteorological practice, the plots start at Coordinated Universal Time (UTC,
or Greenwich time) 00:00. This structure of the covariance matrix follows the behavior of
metabolism of crops. The temporal covariance functions of upwind tower and downwind tower
(not shown in the gure) are very similar, which makes drawing conclusion based on the com-
parison between them becomes dicult.
47
Figure 3.18 Mean components of the CO
2
concentrations at upwind and downwind towers.
Figure 3.19 Temporal covariance of CO
2
surface concentration
48
Eigenvalues and cumulative fraction of the rst several modes are plotted in Fig. 3.20.
Based on Eqn. 3.8, the cumulative of the rst M eigenvalues account for the variance of the
CO
2
concentration eld. It can be seen from the gure that the rst eigenvalue accounts for
about 99.6% of the total variance of the eld, which could imply the CO
2
concentration eld
has a very simple structure so that it has only one dominating mode. By simply compare the
rst eigenvalue (since it is so prominent) of upwind and downwind data, we found that the
variance of CO
2
concentration at downwind position decreases by 8.5%.
Figure 3.20 Eigenvalues (left) and cumulative fraction (right) of the rst 20 modes.
The rst temporal mode of the upwind and downwind eld are given in Fig. 3.21. This
gure shows the pattern of uctuation of CO
2
concentration through time.
To sum up, following primary results are found in this analysis:
1. The mean CO
2
concentration at downwind tower increases by 1.08%
2. The variance of CO
2
concentration at downwind tower decreases by 8.5%
3. The CO
2
concentration eld has only one dominating mode
These results suggest that wind turbine may have very limited eect on the crops growth
with respect to CO
2
concentration which is regarded as to be more crucial to crops growth
than CO
2
ux. On the other hand, CO
2
ux is more likely to be aected by wind turbine since
49
Figure 3.21 The rst temporal mode of CO
2
surface concentration
surface ux consists information of turbulence that will certainly agitated by the wind turbine
rotor. Therefore, analysis on CO
2
surface ux could show more interesting results.
3.3 Numerical Example 3: Full-eld Stochastic Wind Simulation
The meteorological data used in previous two examples have limitations. The CWEX-11
data measures the wind prole at two locations: 4.5 meters and 10 meters. Obviously, this
data do not describe the wind speed prole at hub height. Similarly, CWEX-10 only measures
CO
2
concentration at height 6.45 m. Measurements at only one or two dierent heights are
certainly not enough to capture all the spatial stochasticity in the wind. Third, the data dose
not contain measurements on transverse direction, which makes getting wind ow snapshots on
the plane of turbine rotor becomes impossible. To circumvent this insuciency in measurement,
wind led snapshots on the plane that is perpendicular to the rotor is used, for which certain
time interval has to be chose to construct snapshots. It is worth noting that the framework is
generally applicable to a variety of meteorology data, and the its applicability should not be
aected by choosing dierent snapshot constructing time intervals.
50
3.3.1 Experiment setup
Ideally, a good measurement should include high frequency data of wind speeds at dierent
locations on the rotor plane. The accuracy of using LIDAR (LIght Detection And Ranging)
to measure wind eld has been investigated in [61]. Result of this research shows a promising
application of LIDAR in wind turbine simulation. Given that more time is needed to put this
technique into practice, a dierent source of data is used in this chapter to generate synthetic
turbulence inow of wind turbine. Specically, a turbulence simulator that is developed by
NREL, named TurbSim [34], is used to generate multiple realizations of a stochastic wind, which
are further used as the source data of the temporal-spatial decomposition (TSD) framework
developed in the research. By doing such analysis, we showcase the ability of TSD to reproduce
the important turbulent statistic properties of given measured wind data.
TurbSim utilizes Veers/Sandia method [68] together with various spectral models, such as
Kaimal [38] model and Von Karman Normal Turbulence Model (NTM) [72] that are recom-
mended in the IEC guidelines [13], to simulate 3-D full-eld turbulence. TurbSim also provides
several spectral modes according to dierent terrain conditions and application scenarios. For
the purpose of illustration, Kaimal spectral model and the IEC exponential coherence model
are used in the simulation. The generated wind is designed for using as inow of NREL 5MW
oshore baseline wind turbine [35]
4
, which has 90 m hub height and 63 m blade tip to rotor
center distance (this simulation is given in chapter 4). Therefore, the wind time-series are
simulated on a 15 15 rectangular grid on the rotor plane that spans a 150 m 150 m area
with the center of the grid located at the center of the rotor, which is shown in Fig. 3.22.
30 ten-minute simulations (realizations) of full eld wind turbulence on this grid were
generated with sampling frequency of 4 Hz. They serve as the original wind data of the
TSD. The wind inow can be looked as a time variant wind speed eld on the rotor plane.
Snapshots of this wind eld are taken every 0.25 second (since the sampling frequency is 4 Hz).
For ten-minute simulations, there are 2400 snapshots available. In the simulation, TurbSim
used a referencing wind speed of 6.0 m/s at 10.0 m height, which results in about 8.3 m/s mean
4
It is noteworthy that the NREL 5MW turbine is a conceptual design, thus no structural and material
specication is provided.
51
150 m
1
5
0

m

9
0

m

p
0
p
1
p
2
p
3

Figure 3.22 15 15 grid on the 150 m 150 m rotor plane (yz plane) of the NREL 5MW
wind turbine.
wind speed at hub-height.
Note that TurbSim simulates all three components of turbulence. The along-wind com-
ponent u has the strongest correlation over dierent spatial scales among the three compo-
nents [56]. For the purpose of demonstration, only the along-wind turbulence component is
decomposed in this analysis.
5
3.3.2 Result: statistical comparison
The temporal covariance function of the TurbSim simulated along-wind turbulence eld
is illustrated in Fig. 3.23. Comparing to Fig. 3.3 in the rst numerical example, this gure
has a more uniform pattern. This is because a 10-minute turbulence is regarded as stationary
whereas a 24-hour wind ow is apparently transient. Fig. 3.24 shows the eigenvalues and the
cumulative fractions of the rst several modes. Since the dimensions of temporal covariance
function are 2400 2400, which is much bigger than the case in numerical example one, much
more number of BD modes are needed to get accurate parameterizations. In addition, the rst
5
It does not mean the other two turbulence components are not important in wind turbine simulation.
52
four eigenvalues are so close that the four corresponding modes are all crucial in contributing
the randomness of the turbulence eld. As seen in Fig. 3.25, the rst spatial mode that accounts
for the largest portion of total turbulence energy is mostly uniform, the second and third modes
describe vertical and lateral shear pattern, and the fourth mode possess a more complicated
pattern.
Figure 3.23 Covariance function of the TurbSim simulated turbulence.
Following the same process described in section 3.1.4, we simulated the synthetic wind
that preserves temporal and spatial correlations of the TurbSim generated full-eld wind. In
Fig. 3.26, reconstructed time-series of the turbulence at the rotor center (using 2, 5, and 10
BD modes) are compared with the TurbSim full-eld wind. Not surprisingly, the more modes
used in the simulation, the greater detail in the original turbulence is preserved.
To determine how many modes to use in order to get reasonably accurate representation,
we compare the PSDs of the synthetic turbulence (using 2, 5, and 10 BD modes) at rotor center
with the PSD of the TurbSim full-eld wind in Fig. 3.27.
In addition, the coherence spectra (calculated using 2, 5, and 10 BD modes) between hub
center p
0
and three lateral positions p
1
, p
2
, and p
3
(see Fig. 3.22), are shown in Fig. 3.28
(a), (b), and (c) respectively. It is seen in the two gures that by using only 5 modes, the
53
(a) (b)
Figure 3.24 Spectrum of covariance matrix, (a) eigenvalues, (b) cumulative fraction of energy.
150



75



0

y
z
w
u
v
1
2
4
3
V
e
r
t
i
c
a
l

p
o
s
i
t
i
o
n
,

m

0
75
150

1
= 30.06, (m/s)
2

2
= 16.50, (m/s)
2

3
= 9.97, (m/s)
2

4
= 5.88, (m/s)
2

Figure 3.25 Mean of the rst four stochastic spatial modes.
54
T
S
D

a
n
d

T
u
r
b
S
i
m

g
e
n
e
r
a
t
e
d

t
u
r
b
u
l
e
n
c
e
,

m
/
s

Time, second
TurbSim
10 modes
5 modes
2 modes
Figure 3.26 Reconstructed time series of the along-wind turbulence at the rotor center com-
pared with the original TurbSim simulated ow.
reconstructed turbulence closely reproduces the original wind. In addition, it seems that the
coherence of further separated locations is preserved better by TSD.
3.4 Discussion and Conclusion
In this chapter, three numerical examples of TSD are given. The results of these exam-
ples show that TSD is promising in representing turbulence. However, it is worth to mention
following issues. First, the meteorological data used in this analysis is measured by only one
met tower. Because the lack of the measurement on transverse direction, getting wind ow
snapshots on the plane of turbine rotor becomes impossible. To circumvent this insuciency
in measurement, wind led snapshots on the plane that is perpendicular to the rotor is used,
for which certain time interval has to be chose to construct snapshots. It is worth noting that
the framework is generally applicable to a variety of meteorology data, and the its applicability
should not be aected by choosing dierent snapshot constructing time intervals. In addition,
the framework is able to incorporate both short-term (10-minute) and long-term (years) tem-
poral coherences as long as corresponding data is available. Third, while the mathematical
55
Figure 3.27 PSDs of the synthetic turbulence at rotor center using 2, 5, and 10 modes
(a) (b) (c)
Figure 3.28 Coherence spectra (calculated using 2, 5, and 10 BD modes) between hub center
p
0
and three lateral positions p
1
, p
2
, and p
3
(see Fig. 3.22).
56
framework developed here is used to analyze wind speed, it can also be used to represent other
atmospheric data such as temperature and carbon dioxide ux. This framework can also be
naturally extended to represent ocean waves, which is crucial for o-shore wind turbine siting,
layout and design analysis.
57
CHAPTER 4. WIND TURBINE SIMULATION BASED ON
STOCHASTIC WIND
In the previous chapter, three numerical examples are given to showcase the ability of TSD
in wind eld modeling. In this chapter, the result of the third numerical example 3.3 is used
to perform stochastic analysis on the NREL 5MW oshore wind turbine model.
In general, stochastic analysis is done by representing the random input, modeling the
system of interest, incorporating the randomness in system model or its inputs, and getting
probability of system output. Therefore, this chapter is organized in the same fashion. In
section 4.1, the wind turbine simulation software (FAST) and the NREL 5MW wind turbine
model are introduced. Section 4.2 briey describes the sparse grid collocation algorithm that
is used in the stochastic analysis. The implementation details are introduced in 4.3. In section
4.4, simulation results are given and discussed. Section 4.5 nally concludes this chapter.
4.1 NREL 5MW Wind Turbine Model and FAST
FAST (Fatigue, Aerodynamics, Structures, and Turbulence) is a wind turbine simulator
developed by NREL (National Renewable Energy Laboratory) [36]. It is able to predicting
the responses of both two- and three-bladed horizontal-axis wind turbines with respect to both
fatigue and extreme loads. In FAST, the wind turbine system is divided into a nite number
of rigid/exural bodies connected with elastic joints. The rigid bodies include the tower base,
nacelle, and hub; the exible bodies are blades, drive shaft, and tower. The behavior of the
system is described by a few ordinary/partial dierential equations.
FAST code contains a aerodynamic subroutine package, called AeroDyn [48], that is used to
generate aerodynamic loads along the blade and tower. A schematic of how FAST operates with
58
AeroDyn is shown in Fig. 4.1. Both FAST and AeroDyn need few input les to operate. Input
les for FAST include a primary input le that contains the data for simulation control, turbine
control, initial and environmental conditions, turbine conguration, and output denition; a
tower input le that describes tower structure and tower mode shapes; and a blade input le
contains blade structure parameters and blade mode shapes. Input les for AeroDyn contain a
primary input le that species simulation conguration, directories of airfoil input les, and
information of blade nodes. During the simulation process, FAST communicates with AeroDyn
so that aerodynamic force along the blade at each step is calculated. Finally, FAST provides
a summary output le about the entire simulation and an output le contains time-series of
output variables that was dened in the FAST primary input le.
FAST AeroDyn
Primary
Tower
Blade(s) Primary
Airfoil(s)
Wind
Summary Time-series
Figure 4.1 Schematic of how FAST operates with AeroDyn and their input/output les.
The wind turbine that is used in the following simulation is NREL 5 MW oshore baseline
wind turbine [35]. According to the specication of NREL 5MW wind turbine, the blade length
is 61 m. Three blades are connected with the hub whose radius is 2 m to form a rotor with
radius equals 63 m. The blade surface is composed of a series of airfoil shapes stacked along
axial direction. Starting from the cylinder blade root, the airfoil shape is smoothly shifted into
a series of DU (Delft University) airfoils and NACA64 airfoils. Airfoil cross-sections that are
used in the model are shown in Fig. 4.2. Table 4.1 shows the geometry denition of the blade
model. Fig. 4.3 provides a graphical demonstration of the specications of the blade geometry.
59
Figure 4.2 Airfoil cross-sections used in the design of the wind turbine rotor blades. [37]
Figure 4.3 Illustration of quantities from Table 4.1. [37]
Table 4.1 Wind turbine rotor geometry denition. [37]
RNodes (m) AeroTwst (deg.) Chord (m) AeroCent (-) AeroOrig (-) Airfoil
2.0000 0.000 3.542 0.2500 0.50 Cylinder
2.8677 0.000 3.542 0.2500 0.50 Cylinder
5.6000 0.000 3.854 0.2218 0.44 Cylinder
8.3333 0.000 4.167 0.1883 0.38 Cylinder
11.7500 13.308 4.557 0.1465 0.30 DU40
15.8500 11.480 4.652 0.1250 0.25 DU35
19.9500 10.162 4.458 0.1250 0.25 DU35
24.0500 9.011 4.249 0.1250 0.25 DU30
28.1500 7.795 4.007 0.1250 0.25 DU25
32.2500 6.544 3.748 0.1250 0.25 DU25
36.3500 5.361 3.502 0.1250 0.25 DU21
40.4500 4.188 3.256 0.1250 0.25 DU21
44.5500 3.125 3.010 0.1250 0.25 NACA64
48.6500 2.310 2.764 0.1250 0.25 NACA64
52.7500 1.526 2.518 0.1250 0.25 NACA64
56.1667 0.863 2.313 0.1250 0.25 NACA64
58.9000 0.370 2.086 0.1250 0.25 NACA64
61.6333 0.106 1.419 0.1250 0.25 NACA64
62.9000 0.000 0.700 0.1250 0.25 NACA64
60
The above wind turbine blade model provides us a deterministic solver B where an one-on-
one relationship between the wind speed input v time-series and the turbine response output
time-series u exists, i.e.
B(u : x, v) = 0. (4.1)
where x T are the coordinates in R
3
that denes the location of the output time-series on
the wind turbine structure.
For a stochastic turbulent input that is generated by TSD, the above equation becomes
B(u : x, v()) = B(u : x, ) = 0. (4.2)
where is a set of TSD generated random variables that represent the randomness in the
synthetic turbulence. If m terms of decomposition are used in BD and n terms in KLE are
used to describe each BD term, the resulting random variables =
i

p
i=1
where p = mn.
Eqn. 4.2 denes a stochastic solver where the output of interest u is the stochastic response
of p-dimensional random inputs at location x, i.e. u(x, ).
4.2 Uncertainty Quantication Using Adaptive Sparse Grid Collocation
Method
In previous section, we introduced a stochastic wind turbine system whose wind input
is full-eld turbulence. By performing multiple simulations of the system, nite number of
realizations of the stochastic output can be calculated. The problem now becomes how to solve
Eqn. 4.2 and nd the approximate solution u(x, ).
Let
N
=
i

M
i=1
be a set of nodes in the N-dimensional random space , where M is the
number of nodes. Consider a smooth function u : R
N
R, the Lagrange interpolation of u
can be written as
Ju(x, ) =
M

k=0
u(x,
k
) L
k
(), (4.3)
where L
k
() are Lagrange polynomials given by
L
k
() =
M

i=1,i=k

i

i
. (4.4)
61
u(x,
k
) is the value of u at the given node
k

N
. Therefore, the value of u at any point
can be approximated by Ju(x, ).
In the same fashion, can be approximated by

=
M

k=1

k
L
k
(). (4.5)
Substituting Eqn. 4.5 into the system governing equation yeilds
B
_
u : x,
M

k=1

k
L
k
()
_
= 0. (4.6)
The collocation method converted the original stochastic problem to M deterministic problems,
i.e.
B(u : x,
i
) = 0, i = 1, , M. (4.7)
The stochastic solution is approximated by
u(x, ) =
M

i=1
u(x,
k
) L
k
(). (4.8)
Choosing appropriate sampling points for Lagrange interpolation is a crucial rst step for
collocation methods. For the simplest one-dimensional problem, the optimal selection is usually
the Gauss quadrature. It is straight forward to use the tensor product of the one-dimensional
nodes to chose nodes in multi-dimensional random spaces. Let u(x, ) be a function in N-
dimensional space that need to be approximated, the tensor product interpolation formula can
be written as
(J
i
1
J
i
N
)(u)(x, )
=
M
1

k
1
=1

M
N

k
1
=1
u(x,
i
1
k
1
, ,
i
N
k
N
) (L
i
1
k
1
L
i
N
k
N
), (4.9)
where M
k
is the number of nodes used in the interpolation in the k
th
dimension, J
i
k
is the
interpolation function in the i
k
direction, and
i
k
km
is the k
th
m
point in the i
k
direction. According
to the equation, M
1
M
N
number of points are needed in the computation. This number
grows quickly as the N increases. For the case that M points are chose at each direction, the
total number of points in N-dimensional space is M
N
. Full tensor product method will become
impractical when dealing with high dimensional stochastic space.
62
To overcome this diculty, one need to nd a method that can both satisfy the interpolation
accuracy and minimize the computational cost, i.e. use as less of points as possible. To this end,
sparse grid method is used. The sparse grid method, based on Smolyak algorithm, reduces the
number of nodes from full tensor product formula. In other words, it is a subset of full tensor
product grids. It uses limited numbers of nodes while satisfy the prescribed computational
accuracy.
Consider the one-dimensional interpolation function
|
M
(f) =
M

k=1
f(
k
)L
k
. (4.10)
Let
i
be the incremental interpolate dened as

i
= |
i
|
i1
, |
0
= 0. (4.11)
Smolyaks interpolation /
q,N
is given by
/
q,N
(f) = /
q1,N
(f) +

|i|=q
(
i
1

i
N
)(f), (4.12)
where N is the stochastic dimension and q N is the interpolation order. i = (i
1
, , i
N
) and
[i[ = i
1
+ +i
N
. i
k
is the level of interpolation along the k
th
direction.
Let
(k)
= x
i

M
i=1
be the set of nodes that is used to interpolate the one-dimensional
function, where k denotes the order of polynomial that is used in the interpolation. To compute
the sparse grid interpolation function, only the function values at the sparse grid are needed,
i.e.
1
q,N
=
_
qN+1|i|q
(
(i
1
)

(i
N
)
) (4.13)
Adaptive sparse grid collocation method is developed by following the dimension adaptive
quadrature approach [19]. The basic idea of the adaptive procedure is to assess the stochastic
dimensions according to the error of interpolation in that dimension. Details of this method
can be found in [16].
4.3 Computational Implementation
The solution procedure of this chapter can be divided into three steps:
63
A subroutine for computing deterministic solutions.
A subroutine for allocating nodes and building interpolation functions.
A subroutine for post-processing operations.
Deterministic solver: In section 3.3, a full-eld turbulence representation is developed
based on TSD and the along-wind turbulence component that was originally simulated by
TurbSim. The newly simulated along-wind component (by TSD) together with the other two
orthogonal components (lateral and vertical) are combined and then used as the input full-eld
turbulence of AeroDyn. This process is illustrated in the schematic in Fig. 4.4. By going
through this detour, we seek to demonstrate the ability of TSD to reproduce given wind data
(or eld measurements when available). After incorporating all models as a compact solver,
the input of the solver becomes realization of a set of TSD generated random variables, and the
output becomes wind turbine response time-series. This provides us the deterministic solver.
TrubSim
AeroDyn
u, v, w
Full-field wind
TSD
u
v, w
Full-field wind
u'
Figure 4.4 Schematic of incorporating TSD into full-eld wind simulation.
Sampling and interpolation: A package named FT-AdaGiO (Fault Tolerant ADAptive
sparse GrId allOcator) in C++ developed by Xie et al. [75] was used to allocate the sampling
points in the stochastic domain. In this analysis, 5 BD modes and 3 KLE modes for each
BD mode were used, which makes a 15-dimensional stochastic domain. We used FT-AdaGiO
to generate coordinates of nodes according to adaptive SGC algorithm and to construct the
probabilistic output. The full-eld turbulences were generated based on random samples and
used as input of FAST.
64
Post-processing: Post-processing of current analysis includes extracting basic statistical
solutions. Turbine responses include out-of-plane blade root bending moment, fore-aft tower
bending moment, and power generation were calculated. Probability density functions of the
maximum of these responses were estimated by using KDE.
4.4 Results
In this section, we show the results of NREL 5MW wind turbine simulation with TSD
generated full-eld turbulence inow. The output of the simulation is dened (in the primary
input le of FAST) as 39 variables of interest of the wind turbine. Examples of these variables
include electrical generation, blade and tower deections, and forces and moments acted on
blade and tower. The results of a single simulation are given rst followed by the results of
stochastic analysis using SGC method.
4.4.1 Time responses
Time responses of power generation by using 2, 5, and 10 modes of BD are shown in Fig. 4.5.
It is seen that all four responses have similar structure and the randomness in the wind has
very limited inuence on the power generation. Note that the rated wind speed for NREL
5MW wind turbine is 11.4 m/s, whereas the mean speed of TurbSim wind at hug height is 8.3
m/s. This explains the reason that the average power generation seen in Fig. 4.5 is much less
than the rated 5 MW capacity.
Fig. 4.6 and 4.7 show the 200 seconds response time-series of apwise and edgewise bending
moment at the half span of blade. Fig. 4.8 shows time response of fore-aft bending moment
at tower base. By visual inspection of these gures, it can be seen that a few TSD modes
can accurately preserve the low frequency components of the turbine response. However, to
reproduce the high frequency response seen in the full-simulation based on TurbSim wind,
larger number of TSD terms are required.
To further investigate how well the PSD simulated wind describe the original TurbSim wind,
comparisons between the PSDs of wind turbine time responses corresponding to reconstructed
wind and original wind are given in Fig. 4.9 and 4.10. It is seen in the two gures that PSD
65
P
o
w
e
r

G
e
n
e
r
a
t
i
o
n
,

k
W

Time, second
Figure 4.5 Time response of power generation by using 2, 5, and 10 modes and TurbSim wind
(top to bottom).
F
l
a
p
w
i
s
e

b
e
n
d
i
n
g

a
t

h
a
l
f

s
p
a
n

o
f

b
l
a
d
e
,

k
N
-
m

Time, second
Figure 4.6 Time response of apwise bending moment at the half span of blade by using 2,
5, and 10 modes and TurbSim wind (top to bottom).
66
E
d
g
e
w
i
s
e

b
e
n
d
i
n
g

a
t

h
a
l
f

s
p
a
n

o
f

b
l
a
d
e
,

k
N
-
m

Time, second
Figure 4.7 Time response of edgewise bending moment at the half span of blade by using 2,
5, and 10 modes and TurbSim wind (top to bottom).
F
o
r
e
-
a
f
t

b
e
n
d
i
n
g

a
t

t
o
w
e
r

b
a
s
e
,

k
N
-
m

Time, second
Figure 4.8 Time response of fore-aft bending moment at tower base by using 2, 5, and 10
modes and TurbSim wind (top to bottom).
67
wind is able to locate frequency modes of wind turbine blade and tower vibrations as well as
the TurbSim wind. As expected, as the number of terms in TSD model increases, its accuracy
of estimating frequency responses of wind turbine increases as well.
Figure 4.9 PSD of apwise bending moment at half span of the blade.
Figure 4.10 PSD of edgewise bending moment at half span of the blade.
68
4.4.2 Convergence analysis
Note that the analysis in last section are based on one simulation. With SGC method,
multiple realizations of the stochastic output of the wind turbine system could be achieved,
which allows us to investigate the system in the statistical point of view. Results of SGC
simulation are shown in this section.
In the simulation, m modes of BD and 3 modes of KLE for each BD mode are used. In
other words, to get one realization of the random wind ow, we need to sample from m 3
random variables. Therefore, the number of random dimensions of the wind input is m3. It
is comparatively hard for a system with large number of random input dimensions to converge.
Thus, convergence checks on the computations of SGC by using m = 1, , 4 BD modes are
performed. Fig. 4.11 shows the result of one such convergence analysis where m = 2. It can be
seen in the gure that as the interpolation level increases from 2 to 6, the dierences between
PDFs of succeeding interpolation levels become indistinguishable. To quantitatively check
the rate of convergence, the mean square errors (MSEs) of the inverse cumulative distribution
functions (ICDFs) of dierent levels with respect to level 6 are shown in Table 4.2. We conclude
that for the case of m = 2, good convergence (MSE < 5%) is achieved at interpolation level 4.
Table 4.2 Normalized MSEs of dierent computational levels with respect to level 6, m = 2.
Computational Level # of samples MSE
2 85 1.0
3 230 0.42
4 434 0.033
5 672 0.015
The adaptive sampling procedure of SGC method is achieved by allocating new random
samples in the next computation level based on the surplus of certain or all output variable(s)
in current computation level. In this analysis, the out-of-plane bending moment at blade root is
considered as the output that determines the sampling procedure. Although only the specied
output has best convergence, the rest of the outputs should converge as well since they are
coupled with each other.
According to Fig. 4.11, using the result of interpolation level 4 provides accurate estimations
of the PDFs of out-of-plane bending moment. To check whether consistent convergences of other
69
Figure 4.11 PDFs of out-of-plane bending moment at blade root result from dierent SGC
computation levels.
output variables have been achieved, similar analysis of convergences are done for edgewise and
apwise bending moments and shown in Fig. 4.12 and Fig. 4.13 respectively. The results of
these analysis verify that the assumption on consistent convergence made in previous context
is valid.
Follow the same procedure, convergence checks for cases m = 2, 3, 4 are performed and the
results are shown in Table 4.3. It is worth to mention that the numbers of sampling points
at the convergent level that are result from adaptive SGC method are much smaller than the
result of non-adaptive approach. This fully illustrates the advantages and necessities of using
adaptive SGC method.
Table 4.3 Number of random samples at convergent level for the case of m = 1, 2, 3, 4.
# of BD modes convergent level # of non-adaptive samples # of adaptive samples
1 3 69 40
2 4 1457 434
3 5 26017 11499
4 5 93489 36158
In order to check how many BD modes are sucient in the context of stochastic analysis,
70
Figure 4.12 PDFs of edgewise bending moment at half-span of blade, m = 2.
Figure 4.13 PDFs of apwise bending moment at half-span of blade, m = 2.
71
another convergence analysis on the simulation results with respect to number of BD modes
is performed. PDFs of the out-of-plane bending moment at blade root resulted from dierent
values of m are shown in Fig. 4.14. It is seen that the curve of m = 4 is very close to the
curve of m = 3, which means by using three spatial modes in BD (see the rst three images in
Fig. 3.25), the resulting synthetic wind is sucient to be used as the stochastic input of this
stochastic analysis. In other words, the fourth spatial mode is not prominent enough to aect
the probabilistic output of wind turbine. This implies that the stochastic performance
of wind turbine is only sensitive to larger turbulent structures that are described
by the rst three spatial modes (mean stream ow, vertical shear, and horizontal
shear).
Figure 4.14 PDFs of out-of-plane bending moment at blade root based on the turbulent in-
ows that are constructed by 1, 2, 3, and 4 BD spatial modes.
72
4.5 Discussion and Conclusion
In this chapter, stochastic simulation of NREL 5 MW wind turbine based on full-eld tur-
bulence was performed. Results showed time and frequency responses of the wind turbine with
TSD generated turbulence input are in accordance with the responses resulted from the original
TurbSim wind. After that, stochastic analysis on the wind turbine system was performed. The
result of this analysis provides valuable information on the probability distribution of structural
vibration, forces, shearing, and bending of many exible components (tower, blade, and main
shaft) of the wind turbine system.
On the other hand, FAST models the wind turbine as rigid/exible bodies connected with
elastic joints. Though it is sucient in simple applications that only require general deections
of bodies, it does not provide enough information on the local loads and detailed deformations
on each body, which is very important information for blade design and fatigue analysis. In
next chapter, we will perform such analysis on a full-scale 3D wind turbine model with rich
structure details.
73
CHAPTER 5. WIND TURBINE SIMULATION BASED ON
DETERMINISTIC WIND AND STOCHASTIC-ISOGEOMETRIC
APPROACH
In previous chapter, we performed stochastic analysis on a simplied wind turbine model
using full-led turbulent inow. As an attempt to achieve more accurate and detailed results on
the deformation and surface stress distributions of wind turbine blades, this chapter focus on
integrating stochastic analysis with a full-scale 3D pre-bent wind turbine rotor model, which
is developed based on isogeometric analysis. An adaptive sparse grid collocation method is
used to account for the stochastic input. Kernel density estimation is used to estimate the
distribution of random variables that are related to the randomness of wind turbine model.
This analysis gives insights into failure probability and critical regions of blade shell surface,
which provide wind turbine developers a good reference to improve the design of wind turbine
blades.
This chapter is organized as follows. The randomness in the composite material in wind
turbine blades is rst represented in section 5.1. In section 5.2, the comprehensive wind turbine
blade model that is based on isogeometric analysis is introduced. Adaptive sparse grid colloca-
tion (SGC) algorithm is used in the stochastic analysis (see Chapter 3). The implementation
details are introduced in 5.3. In section 5.4, simulation results are given and discussed. Section
5.5 nally concludes the paper.
5.1 Representing Randomness in Wind Turbine Blade Material
There is a trade-o between blade eciency and strength in wind turbine blade design. From
the aerodynamic point of view, the ideal blade should be as thin as possible. It also needs to
74
be light to minimize loads on the nacelle and tower, and to minimize the manufacturing cost.
On the other hand, blade must be strong enough to withstand the bending load that is caused
by lift and gravity. As a result, existing wind turbine blades have a common structure which
balances their performances with regards to these aspects. The cross section of a typical blade
structure is shown in Fig. 5.1, where the blade has a hollow structure with thick root and
thin tip. In the structure, spar caps carry the major apwise bending moment; the blade shell
maintains the aerodynamic shape, carries the edgewise bending load, and transfers this load to
spars and other structure components; shear web joins the two spar caps and withstands the
shearing load caused by tension and compression on both sides.
Spar caps
Shear webs
Blade shell
Structural cores
Figure 5.1 Wind turbine blade cross section.
The material used in wind turbine blades are usually ber-reinforced composite material.
This is because the material has larger strength-weight ratio than wood and metal and is
also much cheaper than high quality carbon ber. It is worth noting that berglass is mostly
unidirectional and thus exhibits strength in one direction. Biaxial fabric (45

) is used to
make the shear web and some part of the shell so that it bears shear stress most eectively.
When it comes to the blade shell, both bers along the diagonal and bers along axial and
radial directions are needed to resist the torsion loads as well as both apwise and edgewise
bending loads. Therefore, the shell is made of laminated sandwich composites which consist of
multiple plies in dierent directions. The choice of ply directions could be triaxial (45

, 0

)
or quadraxial (45

, 0

, 90

) where material in certain direction takes up to certain percentage


of the space. For example, one such shell layup could be 16 plies with 90% of the ber oriented
in the 0

/90

direction and 10% in the 45

direction. Manufacturers may have dierent


ply layup procedures. It is noteworthy that the plies that are used in the layup process are
usually smaller pieces compared to the length of the blade. To cover the entire blade geometry,
75
multiple plies are needed on each layer of fabric, which results in dierent concentration of ber
toes at dierent locations. In addition, ber alignments are subject to variabilities due to the
prevalence of manual layup processes.
Resin infusion processes are performed after all plies are properly placed on the blade mold.
In general, there are two types of resin infusion processes [22]. Resin transfer molding (RTM) is
one of the commonly used methods. Under pressure, low viscosity resin is pushed or pulled (by
pressure dierence between atmosphere and vacuum) into the mold where several glass ber
cloths are placed. The other method is resin lm infusion (RFI), where partially cured resin
lms are stacked with berglass textiles in a mold. Followed by a process that applies pressure,
heat and vacuum, the stack is fully cured such that resin is infused throughout the mold. Both
processes have to be controlled carefully to prevent occurrence of defects (void, waviness).
Due to this random nature of wind turbine blade manufacturing process, uncertain material
properties inevitably exist between blades. To represent this randomness, the properties that
are required to fully describe the material need to be identied rst.
Fiberglass laminate is an orthotropic material. Since wind turbine blade shell mainly carry
loads on edgewise and apwise directions, the material properties along the thickness are not
signicant. In this case, 4 properties are used to describe orthotropic blade shell: elastic
modulus on apwise (axial) direction E
1
, elastic modulus on edgewise (transverse) direction E
2
,
shear modulus G
12
, and major Poissons ratio
12
. Based on previous discussion, 4 properties
are actually random variables that fully describe the stochastic behavior of composite material
in wind turbine blades. We use the composite materials handbook (MIL-HDBK-17-2F) [65] as
a guide in getting properties of composite materials. The probability density functions of all
material properties are assumed to be Gaussian. In addition, the handbook suggests following
relationship that denes
12
,

12
=

t
1
, (5.1)
where
t
1
and
2
denote axial tension strain and transverse strain respectively. The probability
density function of random variable
12
is estimated using an algorithm that is developed by
Marsaglia [45]. Finally, the mean and standard deviation of all 4 properties of the berglass
76
that is used in this research are listed in Table 5.1
1
.
Table 5.1 Properties of berglass that is used.
(Pa) (Pa)
E
1
46.54 10
9
2.03 10
9
E
2
12.69 10
9
0.26 10
9
G
12
4.69 10
9
0.25 10
9

12
0.13 0.01
5.2 Wind Turbine Blade Model Based on Isogeometric Analysis
The primary advantage of isogeometric analysis is the solution space for dependent variables
can be represented in terms of the same functions that represent the geometry. The isogeometric
procedure that is used in our wind turbine rotor analysis is based on Non-Uniform Rational
B-Splines(NURBS).
Just as its name implies, NURBS are developed from B-splines. Unlike FEM whose physical
domain is divided into sub-domains, B-spline parametric space is divided into patches. Dene
a set of coordinates in one dimension of the parametric space as a knot vector, denote by
=
1
,
2
, . . . ,
n+p+1
, where
i
R, i = 1, 2, . . . , n+p +1 are the knots, p is the polynomial
order, and n is the number of basis functions. Note that p = 0, 1, 2, 3 represents constant,
linear, quadratic, and cubic piecewise polynomials respectively. If knots are equally spaced in
the parametric space, they are said to be uniform. Correspondingly, non-uniform knots are
unequally spaced.
B-spline basis functions are dened recursively as
N
i,p
() =

i

i+p

i
N
i,p1
() +

i+p+1

i+p+1

i+1
N
i+1,p1
(). (5.2)
where p = 1, 2, 3, . . ., and
N
i,0
() =
_

_
1 if
i
<
i+1
,
0 otherwise.
(5.3)
1
The material is S2-449 17k/sp 381 unidirectional tape whose resin content is 28-29 wt%; ber volume is
50.1-54.0 %; density is 1.85-1.92 g/cm
3
; and ply thickness is 0.0033-0.0037 in.
77
With this denition, one may notice that for p = 0 and 1, B-spline basis functions are the same
standard piecewise constant and linear basis functions in nite element method. However, this
identity does not hold for p 2. Repeated knots are allowed in knot vectors. A knot vector
is said to be open if its rst and last knots appear p + 1 times. Basis functions over open
knot vectors are interpolatory at the ends of the parametric space interval, [
1
,
n+p+1
], and
generally not interpolatory at knots within the interval. In general, basis functions in order p
are C
p1
-continuous. The continuity at a knot decreases by k if the knot is repeated k times.
The basis function is interpolatory at a knot if it is repeated p times. Finally, it is worth noting
that the support of N
i,p
is [
i
,
i+p+1
] and

n
i=1
N
i,p
() = 1 for any .
Now that B-spline basis function is dened, B-spline curves in R
d
, denoted by C(), can be
constructed by taking linear combination of B-spline basis functions
C() =
n

i=1
N
i,p
()B
i
. (5.4)
where B
i
R
d
, i = 1, 2, . . . , n, are referred to as control points.
Consider a control net with control points B
i,j
, i = 1, 2, . . . , n, j = 1, 2, . . . , m, and knot
vectors =
1
,
2
, . . . ,
n+p+1
, and 1 =
1
,
2
, . . . ,
m+q+1
, a B-spline surface can be
dened as
S(, ) =
n

i=1
m

j=1
N
i,p
()M
j,q
()B
i,j
, (5.5)
where N
i,p
and M
j,q
are p
th
order and q
th
order B-spline basis functions respectively.
Similarly, given control net B
i,j,k
, i = 1, 2, . . . , n, j = 1, 2, . . . , m, k = 1, 2, . . . , l, and
knot vectors =
1
,
2
, . . . ,
n+p+1
, 1 =
1
,
2
, . . . ,
m+q+1
, and : =
1
,
2
, . . . ,
l+r+1
,
a B-spline solid is dened by
S(, , ) =
n

i=1
m

j=1
l

k=1
N
i,p
()M
j,q
()L
k,r
()B
i,j,k
, (5.6)
where L
k,r
are r
th
order B-spline basis functions.
To obtain exact geometry in R
d
, projective transformation of B-spline entity in R
d+1
must
be done. Let B
w
i
be a control net for a B-spline curve in R
d+1
with knot vector . Control
points for desired NURBS curve in R
d
are derived from B
w
i
which are referred as projective
78
control points for the desired curve.
(B
i
)
j
= (B
w
i
)
j
/w
i
, j = 1, . . . , d,
w
i
= (B
w
i
)
d+1
. (5.7)
where (B
i
)
j
is the j
th
component of vector B
i
, and w
i
is referred to as the i
th
weight. Non-
Uniform Rational B-Splines(NURBS) curve is given by
C() =
n

i=1
R
p
i
()B
i
, (5.8)
where R
p
i
() are rational basis functions given by
R
p
i
() =
N
i,p
()w
i

n
k=1
N
k,p
()w
k
, (5.9)
Rational surfaces and solids could be dened similarly
R
p,q
i,j
(, ) =
N
i,p
()M
j,q
()w
i,j

i=1

j=1
N

i,p
()M

j,q
()w

i,

j
. (5.10)
R
p,q,r
i,j,k
(, , ) =
N
i,p
()M
j,q
()L
k,r
()w
i,j,k

i=1

j=1

k
N

i,p
()M

j,q
()L

k,r
()w

i,

j,

k
. (5.11)
Isogeometric analysis represents the geometry, motions of the model, properties such as
strain states and etc. with the same basis functions. Based on the fact that NURBS forms
almost the exact geometry of the model, resulting quantities of interest could be more accurate
than the results by using FEA. Therefore, the rotor and surrounding uid are modeled by
NURBS in this work.
The NREL 5MW wind turbine blade model is constructed by using isogeometric analysis,
which is shown in Fig. 5.2. The surrounding uid domain is modeled with volumetric NURBS
so that we can analyze the FSI between them. It is worth noting that the rotor can be divided
into three equal portions, thus only one third of the rotor and the uid domain is modeled
for the purpose of increasing computational eciency. The dimensions of the entire problem
domain and the NURBS mesh are shown in Fig. 5.3. More details of modeling wind turbine
rotor based on isogeometric analysis can be found in [7, 8].
The above wind turbine blade model provides us a deterministic solver where an one-to-one
relationship between input and desired output exists. In our stochastic analysis, the inputs are
79
Figure 5.2 Surface meshes of the NREL 5MW wind turbine blade. . [9]
(a) (b)
Figure 5.3 (a) Volumetric NURBS mesh of the computational domain and (b) A planar cut
to illustrate mesh grading toward the rotor blade. [9]
80
the 4 wind turbine material parameters (E
1
, E
2
, G
12
, and
12
), and the outputs of interest are
blade deformation, stress distribution, and etc. i.e.
B(u : x, q) = 0. (5.12)
where B represents the deterministic solver, x T are the coordinates in R
3
, q = q
1
, q
2
, ...
is a set that represents boundary conditions, material properties as well as any condition that
denes the physical system, u is the output of interest at location x. For a deterministic wind
turbine model with no random material properties, one can nd the value of u at any point x
with given q. This results in a function u : T R.
Suppose one of the conditions, q
i
q, denote by for the sake of convenience, is actually
uncertain. Let

be the sample space of which contains possible values that can take,
i.e.

. Dene a -algebra, denote by T, a non-empty set of events in

, such that it
is closed under complements and countable unions. Let T : T [0, 1] be a function returning
the probability of particular event in T happens. According to the properties of -algebra, T
should be countably additive and satisfy that T(

) = 1. The three denitions introduced


above can be composed as a probability space (

, T, T), where T is the -algebra over sample


space

, T is the probability measure on T. Based on above discussion, the deterministic


system has been changed to stochastic system with random input (

) and random output.


It is worth noting that (

) is an abstract quantity that lies in abstract probability space


(

, T, T). The problem becomes


B(u : x, q
i
, (

)) = 0. (5.13)
where q
i
= q
1
, q
2
, ..., q
i1
, q
i+1
, .... The solution of above set of dierential equations is
actually a function u : T

R.
Since (

) is just an abstract concept which could not be quantied. To solve equation


(5.13), the randomness in input must be represented by numerical mean. To this end, dene
a measurable space (R, ), on which lies a real-valued random variable

R. By
measurable, it means for every subset B , its pre-image
1

(B) T where
1

(B) =

) B. With this denition, we relate the abstract quantity (

) to a real-valued
81
quantity

. The solver becomes


B(u : x,

) = 0. (5.14)
where

is the random variable which represents the randomness in the input parameter .
Furthermore, if multiple input parameters are actually random, the above equation is converted
to
B(u : x,
1
,
2
, ...,
k
) = 0. (5.15)
where
i
, i = 1, 2, ..., k are random inputs. Since 4 material parameters are considered to be
random as discussed previously, the wind turbine solver becomes
B(u : x, ) = 0, (5.16)
where =
i

4
i=1
. As a result, the output of interest u is the stochastic response of 4 random
inputs.
The next step of the analysis is to use SGC method to nd the solution u(x, ) of the
stochastic solver (Eqn. 5.15). Mathematical background can be found in section 4.2.
5.3 Computational Implementation
As discussed in section 1.1.3, SGC is used to perform stochastic analysis. It is noteworthy
that SGC assumes all random variables are uncorrelated to each other. However, no analysis
or experiment suggests that the four material properties satisfy this assumption. Yet for the
purpose of demonstration, in this research we assume they are uncorrelated without loss of
generality. In case of correlated random variables, SGC scheme can be used after principal
component analysis (PCA), which is a technique to convert a set of possibly correlated variables
into a set of uncorrelated variables (principal components) [33].
The solution procedure of this chapter can be divided into three steps:
A subroutine for computing deterministic solutions.
A subroutine for allocating nodes and building interpolation functions.
A subroutine for post-processing operations.
82
Deterministic solver: In this research, we are interested in assessing the impact of the
randomness which is caused by the random nature of wind turbine manufacturing process to
rotor performance. To this end, a wind turbine blade model based on isogeometric analysis was
used in the simulation. The blade shell was made of a symmetric berglass-epoxy composite
with [45/0/90
2
/0
3
]
s
layup. The model was then simulated at prescribed steady wind velocity
of 11.4 m/s and rotor angular velocity of 12.1 rpm
2
. Fig. 5.4 shows the simulation setup.
Results of the simulation including the blade wake velocity eld, pressure contour on the blade
surface, blade deformation, shell stress distribution, and etc. More details about the simulation
could be found in [7, 8]. In our analysis, stress distribution on the blade was calculated and
regarded as stochastic output of the system.
Figure 5.4 Simulation setup.
Sampling and interpolation: FT-AdaGiO was used to allocate the sampling points in
the 4-dimensional random domain. The package generated coordinates of nodes
i
=
j

[0, 1], j = 1 4 according to adaptive SGC algorithm and construct the probabilistic output.
The random material properties of the blade were generated based on
i
and incorporated
in the deterministic solver. After that, outputs were calculated based on the random material
properties. Although the number of samples greatly reduced by using adaptive SGC algorithm,
solving the complicated deterministic solver so many times is still a time consuming task.
Post-processing: Post-processing of current analysis includes extracting basic statistical
2
Ideally, stochastic inow should have been used, however, the wind turbine model used in this experiment
does not allow us to implement stochastic wind in the simulation. Thus, steady wind was used.
83
solutions. Mean and standard deviation of stress distribution on blade shell can be calculated
with all realizations of the output of interest. Probability density function of the stress value
at critical location on the blade can also be generated by KDE.
5.4 Results
The calculation of adaptive SGC procedure was carried out up to a level of interpolation
(6) which has 168 sampling points. In order to make sure the selected samples are appropriate
to get accurate solutions, the convergence property of the SGC framework was checked by
comparing the PDFs at a critical location (where comparatively large stress occurs) of dierent
SGC computation levels. The result of the convergence check is shown in Fig. 5.5. From
the gure we notice that the dierence between two consecutive levels decreases drastically as
interpolation level increases, which implies good convergence. To have a better view of the
convergence, the mean square error (MSE) of inverse cumulative distribution functions (ICDF)
of dierent levels with respect to level 6 are shown in Table 5.2. Good convergence of the
approximation error that is shown in Table 5.2 indicates using results of level 6 is sucient.
Stress (MPa)
P
D
F
64 66 68 70 72 74
5E-08
1E-07
1.5E-07
2E-07
Level2
Level3
Level4
Level5
Level6
Figure 5.5 PDFs at a critical location of dierent SGC computation levels.
Fig. 5.6 (a) shows one of the 3D scatter plots of the nodes that are allocated in computation
84
Table 5.2 Normalized MSE of dierent computational levels with respect to level 6.
Computational Level # of samples MSE
2 8 1.0
3 40 0.28
4 88 0.064
5 136 0.0068
level 6. It can be seen from the gure that the nodes are sparsely distributed in the stochastic
domain. As discussed previously, the number of samples used in computation level six is 168.
The advantage of the adaptive procedure can be seen clearly by comparing Fig. 5.6 (a) with
Fig. 5.6 (b) where 1824 samples are generated according to non-adaptive SGC algorithm.
Although much less sampling points was used in the adaptive method, good convergence was
achieved (see Fig. 5.5). Fig. 5.6 (a) also sheds light on the physical properties of the system.
For example, the plot shows a inside-sparse structure which means taking more samples from
the region where E
1
, E
2
, and G
12
are in their intermediate values does not help a lot in rening
the stochastic results.
E
1
0
0.2
0.4
0.6
0.8
1
E
2
0
0.2
0.4
0.6
0.8
1
G
1
2
0
0.2
0.4
0.6
0.8
1
E
2
0
0.2
0.4
0.6
0.8
1
E
1
0
0.2
0.4
0.6
0.8
1
G
1
2
0
0.2
0.4
0.6
0.8
1
(a) (b)
Figure 5.6 Sampling points allocated by adaptive SGC algorithm (a) and non-adaptive SGC
algorithm (b) at computation level 6.
In order to increase the clearance between blade tip and turbine tower, large turbine blades
are pre-bent such that the clearance is still sucient after wind load is applied and blades are
deformed. Blade deformation is shown in Fig. 5.7. It can be seen in the gure that after the
deformation, the blade is almost straight. Noting that the 11.4 m/s wind speed is close to the
85
rated wind speed of most wind turbines, this result implies that pre-bent rotor will have the
optimized power output at the rated wind speed. Similar analysis can be done to become a
good reference of designing the optimal curvature of pre-bent blades.
11.4 m/s
Deformed shape
Pre-bent shape
Figure 5.7 Blade deformation under 11.4 m/s wind load and 12.1 rpm rotating speed.
5.4.1 Stress analysis
The deformation of the blade causes stress on the blade structure. A blade design has
to be veried such that the stress introduced by structure deformation will not exceed the
design strength of the material. According to the GL design code [18], the design strength
is the characteristic value of the material divided by the partial safety factor of the material,
where the characteristic value is calculated based on experiment results and corresponds to
95% survival probability and condence level.
Since the apwise bending load dominates the deformation of blades, mean contour plot of
stress along apwise direction on the 13
th
ply, where maximum mean stress exists, is shown in
Fig. 5.8. The red region on the Fig. 5.8 shows the area that suers the largest average apwise
stress over the entire blade surface.
Although this gure gives us a general idea of the critical region, it does not provide all the
information of the statistical behavior of the stress. A point with low mean and high variance
of stress could be more critical than the point that has the maximum mean stress because high
86
Stress, Pa
Figure 5.8 Mean stress contour the 13
th
ply.
variance may result in unexpected peak of stress that may damage the blade. Therefore, a
plot of standard deviation of apwise stress on the 13
th
ply is shown in Fig. 5.9 (Noting that
the 13
th
ply is where maximum standard deviation of apwise stress exists). Comparing with
Fig. 5.8, the point with maximum standard deviation is almost identical to the point with
maximum mean.
Stress, Pa
Figure 5.9 Contour of the standard deviation of stress on the 13
th
ply.
We are interested in the probability density function at critical region. KDE method (see
87
Appendix 2.1.3) was used in estimating PDFs. Fig. 5.10 shows the PDF of the node that has
the maximum standard deviation of stress values. It is interesting to note that the PDF is
Gaussian-like that is commonly used for modeling physical phenomena.
Stress, Pa
x
P
D
F
6.5E+ 07 7E+ 07 7.5E+ 07
5E-08
1E-07
1.5E-07
2E-07
Figure 5.10 PDF of the stress at the point that has the maximum standard deviation of
apwise stress.
According to the GL design code [18], the design strength R
d
= R
k
/
Mx
. In this equation,
R
k
is the characteristic value given as
R
k
(, n) = x
_
1
_
1.645 +
1.645

n
__
, (5.17)
where x denotes the mean of the test strength and n is the number of test values.
Mx
can be
calculated as follows

Mx
= 1.35

i
C
ix
, (5.18)
where C
ix
is the reduction factors decided by dierent types of analysis (for material produced
by resin infusion method, and for short-term strength verication,
Mx
= 1.35 1.1 = 1.485).
According to the test method in [65], the characteristic value of the material that is used
in current analysis is 1.68 GPa. As a result, the design strength of the given material is
R
d
= 1.68/1.485 = 1.13 GPa, which means any stress on the blade that is higher than R
d
would fail the stress verication. In Fig. 5.10, the apwise stress takes values that are much
88
smaller than the design strength of the material, which means current design is safe but either
the blade material is unnecessarily strong or the structure design is redundant.
With probability density function at critical points, we can further investigate the proba-
bility of the stress at a critical point exceeding the design strength of the material. For the
purpose of demonstration, consider the case whose design strength R
d
= 68MPa. It is valuable
to know where on the blade would break rst. According to the denition of probability density
function, the area for stress value > R
d
on the PDF equals the probability of the point has
stress exceeds the design tensile strength of the material. Critical regions that have a large
probability of failure can be calculated. As shown in Fig. 5.11, value at each point represents
the probability of the stress at that point exceeds the design tensile strength. It can be seen
in Fig. 5.11 that critical regions are located at the root transition area and half-length of the
blade.
Failure Probability
Figure 5.11 Critical region that could have stress higher than 68MPa with large probability.
Remark 3. The current example only shows the result of the laminate failure analysis in
normal load case. However, the framework presented in this research can be used in many
other analysis which includes analysis on laminate failure and stability failure with regard to
the short-term strength, fatigue strength and stability in all load cases (normal/operational and
extreme) [18].
89
5.4.2 Failure analysis
In the previous section, the analysis on apwise stress was provided. However, failure in
orthotropic material depends on stress components in all axis and the interaction between them.
Tsai-Wu failure criterion [63] is widely used for identifying failures in the analysis of anisotropic
composite materials. In Tsai-Wu criterion, failure is evaluated based on a combination of
stress components. It is designed such that the combination of dierent stress components are
considered for failure initiation in composite materials. The failure index can be calculated
based on a multi-axial stress condition and regarded as an important indicator of failure.
A failure index that is greater than 1.0 means failure occurs.
3
Contour plot of the mean
and standard deviation of failure index (calculated based on the actual design strength R
d
=
1.13 GPa) on the blade surface was shown in Fig. 5.12. As seen in Fig. 5.12 (a), the maximum
mean value of failure index ( 0.4) is much smaller than 1.0. Fig. 5.12 (b) shows the standard
deviations of failure index. In addition, it is noteworthy that maximum failure index is located
at the 14
th
ply rather than the 13
th
ply that contains maximum apwise stress. Although
falpwise stress is a very important indicator in wind turbine structure analysis, it is not enough
to predict failure.
To further investigate the probability that the maximum failure index exceeds 1.0, the
PDF of failure index at the location that has maximum mean failure index on the blade was
estimated and is given in Fig. 5.13. Result shows that the probability of failure (FI > 1.0) is
nearly zero. This analysis provides us a unique method to check wether a wind turbine design
is valid. We envision this analysis as a useful reference in wind turbine design.
5.5 Discussion and Conclusion
Current wind turbine simulations are mostly based on simplied blade models which could
not meet the requirement for accurately analyzing the uid structure interaction between blade
and surrounding air domain. On the other hand, in spite of the fact that randomness inher-
ent in blade manufacturing process may impact blade performance and turbine eciency, the
3
Tsai-Wu criterion does not provide the information about when and how the failure occurs.
90
(a)
(b)
Tsai-Wu FI
Tsai-Wu FI
Figure 5.12 Contour of failure index on the 14
th
ply.
91
Failure Index
P
D
F
0.4 0.5 0.6 0.7 0.8 0.9 1
5
10
15
20
25
30
Figure 5.13 Probability density function of failure index at the point that has maximum mean
failure index on the blade.
mechanism of randomness aecting the performance of wind turbines is still an open question.
In this research, we try to solve these problems by integrating stochastic analysis, which is
based on adaptive sparse grid collocation method, on a full scale pre-bent 3D wind turbine
rotor model constructed with isogeometric analysis. With a given steady wind load, the model
gives the deformation and the stress distribution of the blades. Four material properties regards
to composite laminate were given as stochastic input of the system. Stochastic results were
generated based on 168 sparsely allocated sampling points.
An example of stress analysis and failure analysis on NREL 5MW wind turbine blade
design was given. The results identify the region that is most likely to have large stress or large
variation of stress. Moreover, the probability density functions of the stress values at critical
regions were calculated. Finally, the probability of the critical region to have stress values lager
than given threshold was calculated. If we dene the threshold according to the design strength
of the material, this probability becomes the probability for the blade design to fail. All above
results should give us a better understanding of how the randomness aect the performance of
the blade.
92
CHAPTER 6. SUMMARY AND DISCUSSION
In this research, we explored the feasibility of incorporating stochastic analysis on wind
turbine design. We rst developed a low-complexity yet realistic data-driven turbulence simu-
lation framework using temporal and spatial decomposition. A synthetic turbulent inow was
generated using this framework and used as the stochastic input of a simplied wind turbine
model. We utilized adaptive sparse grid collocation method and performed stochastic analysis
on this system, where primary results on stochastic wind turbines performance were found. As
an attempt to achieve more accurate and detailed results on the deformation and surface stress
distributions of wind turbine blades, a 3D comprehensive wind turbine model that was devel-
oped based on isogeometric analysis and a deterministic wind input were used in simulations.
Stochastic analysis on random composite material properties was done on this comprehensive
wind turbine model, where stress and failure analysis were performed.
From the analysis performed in this research, we found that the temporal spatial decompo-
sition framework is able to precisely reproduce the temporal and spatial statistics of any given
large data set of wind. The random variables that are generated in the decomposition process
showed that Gaussian randomness assumption may not be valid for turbulences in various en-
vironmental conditions. The temporal and spatial modes resulted from the decomposition also
showed valuable information about the turbulence. Simulation results of using TSD synthetic
inow showed good consistency with the results of using original wind data. The results of
stochastic analysis show that the stochastic performances of wind turbines are only sensitive
to few largest spatial modes in the turbulent inow. From the stochastic analysis on the 3D
comprehensive wind turbine model, critical regions on blades were located, the probability of
blades to have failures were calculated, and suggestions of blades design were made.
A data-driven low-complexity model that encodes the spatial- and temporal-correlations
93
and is location- and topography- sensitive has signicant utility in a variety of wind energy
applications. These include, wind farm layout optimization using realistic wind models; robust
control of turbine operations based on the low-complexity stochastic wind model; stochastic
analysis for robust (and lean) design of turbine-blades and other components; and using re-
alistic wind models to analyze wind farm output integration with the power grid. While the
mathematical framework developed here is used to analyze wind speed, it can also be used to
represent other atmospheric data such as temperature and carbon dioxide ux. This frame-
work can also be naturally extended to represent ocean waves, which is crucial for o-shore
wind turbine siting, layout and design analysis. On the other hand, stochastic analysis of wind
turbine provides us a better understanding of how the randomness aect the performance of
wind turbine, which could be a good reference for the wind turbine design guideline.
It is worth noting that although TSD is capable of representing any given wind data, and its
accuracy does not strongly depend on the number of modes used in the decomposition (see the
discussion in Sec. 3.1.5), more modes are still desired for accurately representing the full-eld
wind. As a result, using it as the stochastic input of wind turbine model results in a very high-
dimensional stochastic problem whose convergence is dicult to achieve. In addition, we only
showed the ability of TSD in modeling the along-wind turbulence component that has apparent
temporal and spatial correlations, more work need to be done to show the usefulness of this
framework on representing the transverse and vertical turbulence components. Moreover, the
development TSD framework is still in the initial stage, thus more work need to be done for
it to be used in industry. Furthermore, the availability of high frequency eld measurements
is the bottleneck of applying TSD in data-driven analysis. Finally, although the 3D wind
turbine model that is used in this research has rich surface geometry, it does not have complete
structure components such as shear webs and structural cores. To better estimate stress and
predict failure, rotor models with complete blade structure are required.
Future works of this research include but not limit to the following topics. First, the
computational framework of TSD needs to be tailored such that it is consistent with IEC
industry standard, and easy to be used in existing wind simulation codes. Second, by comparing
the TSD parameterizations of wind turbines inow and wake, possible conclusions on how the
94
statistics structure of turbulence is changed by wind turbines could be drawn. Third, other
properties of blades can be used as stochastic input. For example, random waviness defects
and stochastic structure parameters. Fourth, incorporating TSD framework in complete wind
turbine models, which would reveal more realistic simulation results.
95
BIBLIOGRAPHY
[1] Puneet Agarwal and Lance Manuel. Simulation of oshore wind turbine response for
long-term extreme load prediction. Engineering structures, 31(10):22362246, 2009.
[2] Anders Ahlstrom. Aeroelastic Simulation of Wind Turbine Dynamics. PhD thesis, Royal
Institute of Technology, Stockholm, Sweden, 2005.
[3] Edward B Arnett, W Brown, Wallace P Erickson, Jenny K Fiedler, Brenda L Hamilton,
Travis H Henry, Aaftab Jain, Gregory D Johnson, Jessica Kerns, Rolf R Koford, et al.
Patterns of bat fatalities at wind energy facilities in north america. The Journal of Wildlife
Management, 72(1):6178, 2008.
[4] I. Babuska, F. Nobile, and R. Tempone. A stochastic collocation method for elliptic partial
dierential equations with random input data. SIAM Journal on Numerical Analysis,
45(3):10051034, 2007.
[5] Satish Balay, Jed Brown, Kris Buschelman, William D. Gropp, Dinesh Kaushik,
Matthew G. Knepley, Lois Curfman McInnes, Barry F. Smith, and Hong Zhang. PETSc
Web page, 2012. http://www.mcs.anl.gov/petsc.
[6] Robert MR Barclay, EF Baerwald, and JC Gruver. Variation in bat and bird fatalities
at wind energy facilities: assessing the eects of rotor size and tower height. Canadian
Journal of Zoology, 85(3):381387, 2007.
[7] Y. Bazilevs, M.-C. Hsu, and I. Akkerman. 3D simulation of wind turbine rotors at full
scale. Part I: geometry modeling and aerodynamics. Methods in Fluids, 65:207235, August
2011.
96
[8] Y. Bazilevs, M.-C. Hsu, and J. Kiendl. 3D simulation of wind turbine rotors at full scale .
Part II : Fluid structure interaction modeling with composite blades. Methods in Fluids,
65:236253, October 2011.
[9] Y Bazilevs, M-C Hsu, and MA Scott. Isogeometric uidstructure interaction analysis
with emphasis on non-matching discretizations, and with application to wind turbines.
Computer Methods in Applied Mechanics and Engineering, 2012.
[10] Ian D Bishop. Determination of thresholds of visual impact: the case of wind turbines.
Environment and Planning B, 29(5):707718, 2002.
[11] L Le Cam. Maximum Likelihood: An Introduction. International Statistical Review /
Revue Internationale de Statistique, 58(2):pp. 153171, 1990.
[12] J. A. Carta, P. Pamrez, and S. Velazquez. A review of wind speed probability distribu-
tions used in wind energy analysis: Case studies in the Canary Islands. Renewable and
Sustainable Energy Reviews, 13(5):933955, June 2009.
[13] International Electrotechnical Committee et al. Iec 61400-1: Wind turbines part 1: Design
requirements. International Electrotechnical Commission, 2005.
[14] D Coomans and D L Massart. Alternative k-nearest neighbour rules in supervised pattern
recognition : Part 1. k-Nearest neighbour classication by using alternative voting rules.
Analytica Chimica Acta, 136(0):1527, 1982.
[15] Lev D Elsgolc. Calculus of variations, volume 19. Courier Dover Publications, 2007.
[16] B Ganapathysubramanian and N Zabaras. Sparse grid collocation schemes for stochastic
natural convection problems. Journal of Computational Physics, 225(1):652685, July
2007.
[17] A Gelman, J B Carlin, H S Stern, and D B Rubin. Bayesian Data Analysis. Chapman
and Hall/CRC, London, 2 edition, 2003.
[18] Germanischer Lloyd. Guideline for the Certication of Wind Turbines. Germanischer
Lloyd, Hamburg, 2010 edition, 2010.
97
[19] Thomas Gerstner and Michael Griebel. Dimension-Adaptive Tensor-Product Quadrature.
Computing, 71:6587, 2003.
[20] R. G. Ghanem and P. D. Spanos. Stochastic Finite Elements: A Spectral Approach. Dover
Publications, 1998.
[21] GL Garrad hassan. Bladed: Wind Turbine Design Software, oct 2012. http://www.gl-
garradhassan.com/en/software/GHBladed.php.
[22] D.A. Grin. Blade System Design Studies Volume I: Composite Technologies for Large
Wind Turbine Blades. Technical report, SAND2002-1879, Sandia National Laboratories,
Albuquerque, NM, July 2002.
[23] Qiang Guo. Link of code used in current paper, 2012.
http://www3.me.iastate.edu/bglab/pages/projects.html.
[24] GWE Council. Global Wind 2010 Report. Global Wind Energy Council, Brussels, Belgium,
2010.
[25] Martin O.L. Hansen and Helge Aagaard Madsen. Review Paper on Wind Turbine Aero-
dynamics. Journal of Fluids Engineering, 133, November 2011.
[26] Vicente Hernandez, Jose E. Roman, and Vicente Vidal. SLEPc: A scalable and exi-
ble toolkit for the solution of eigenvalue problems. ACM Transactions on Mathematical
Software, 31(3):351362, 2005.
[27] Philip Holmes, John L. Lumley, and Gahl Berkooz. Turbulence, Coherent Structures,
Dynamical Systems and Symmetry. Cambridge University Press, 1998.
[28] T J R Hughes, J A Cottrell, and Y Bazilevs. Isogeometric analysis: CAD, nite elements,
NURBS, exact geometry, and mesh renement. Computer Methods in Applied Mechanics
and Engineering, 194(39-41):41354195, 2005.
[29] J. Hunter and B. Nachtergaele. Applied Analysis. World Scientic Publishing Company,
Singapore, 2005.
98
[30] X. Jiang, N.-C. Lau, I. M. Held, and J. J. Ploshay. Mechanisms of the Great Plains Low-
Level Jet as Simulated in an AGCM. Journal of the Atmospheric Sciences, 64(2):532547,
February 2007.
[31] Javier Jimenez. Turbulent velocity uctuations need not be gaussian. Journal of Fluid
Mechanics, 376(1):139147, 1998.
[32] Arthur Jobert, Pia Laborgne, and Solveig Mimler. Local acceptance of wind energy:
Factors of success identied in french and german case studies. Energy policy, 35(5):2751
2760, 2007.
[33] I.T. Jollie. Principal Component Analysis. Springer Series in Statistics. Springer, NY,
2nd edition, 2002.
[34] Bonnie J Jonkman. TurbSim users guide: Version 1.50. National Renewable Energy
Laboratory Colorado, 2009.
[35] J Jonkman, S Buttereld, W Musial, and G Scott. Denition of a 5-mw reference wind
turbine for oshore system development. Technical Report NREL/TP-500-38060, (Febru-
ary):75, 2009.
[36] Jason Jonkman. FAST: An Aeroelastic Design Code for Horizontal Axis Wind Turbines,
mar 2012. http://wind.nrel.gov/designcodes/simulators/fast/.
[37] Jason Mark Jonkman, S Buttereld, W Musial, and G Scott. Denition of a 5-MW refer-
ence wind turbine for oshore system development. National Renewable Energy Laboratory
Colorado, 2009.
[38] JC Kaimal, JCj Wyngaard, Y Izumi, and OR Cote. Spectral characteristics of surface-
layer turbulence. Quarterly Journal of the Royal Meteorological Society, 98(417):563589,
1972.
[39] ND Kelley, BJ Jonkman, GN Scott, JT Bialasiewicz, and LS Redmond. The impact of
coherent turbulence on wind turbine aeroelastic response and its simulation. In Windpower
2005 Conference Proceedings, 2005.
99
[40] Y. W. Kwon, D. H. Allen, and R. Talreja, editors. Multiscale Modeling and Simulation of
Composite Materials and Structures. Springer, New York, 2008.
[41] E. Lantz, M. Hand, and R. Wiser. WREF 2012: The Past and Future Cost of Wind
Energy. 2012 World Renewable Energy Forum, May 2012.
[42] Jesper Winther Larsen, R Iwankiewicz, and Sren RK Nielsen. Nonlinear stochastic sta-
bility analysis of wind turbine wings by monte carlo simulations. Probabilistic engineering
mechanics, 22(2):181193, 2007.
[43] D. Lee, D. H. Hodges, and M. J. Patil. Multi-exible-body Dynamic Analysis of Horizontal
Axis Wind Turbines. Wind Energy, (5):281300, 2005.
[44] Jakob Mann. Wind eld simulation. Probabilistic engineering mechanics, 13(4):269282,
1998.
[45] George Marsaglia. Ratios of Normal Variables. Journal of Statistical Software, 16(4), may
2006.
[46] L. Mathelin and O. L. Maitre. Robust control of uncertain cylinder wake ows based on
robust reduced order models. Computers & Fluids, 38(6):11681182, 2009.
[47] A. Messac, S. Chowdhury, and J. Zhang. Characterizing and Mitigating the Wind
Resource-based Uncertainty in Farm Performance. Journal of Turbulence, 13(13):126,
2012.
[48] Patrick J Moriarty and A Craig Hansen. AeroDyn theory manual. National Renewable
Energy Laboratory Golden, Colorado, USA, 2005.
[49] Nicola Barberis Negra, Ole Holmstrm, Birgitte Bak-Jensen, and Poul Srensen. Model
of a synthetic wind speed time series generator. Wind Energy, 11(2):193209, 2008.
[50] Patrik Passon and Martin K uhn. State-of-the-art and Development Needs of Simulation
Codes for Oshore Wind Turbines. In Copenhagen Oshore Conference 2005, Copenhagen,
oct 2005.
100
[51] D. C. Quarton. The Evolution of Wind Turbine Design Analysis-A Twenty Year Progress
Review. Wind Energy, 1(S1):524, 1998.
[52] D. A. Rajewski, E. S. Takle, J. K. Lundquist, S. Oncle, J. H. Prueger, T. W. Horst, M. E.
Rhodes, R. Pfeier, J. L. Hateld, K. K. Spoth, and R. K. Doorenbos. CWEX: Crop/Wind-
energy EXperiment: Observations of surface-layer, boundary-layer and mesoscale interac-
tions with a wind farm. http://journals.ametsoc.org/doi/pdf/10.1175/BAMS-D-11-00240,
2012.
[53] A.L. Rogers, J.F. Manwell, and S. Wright. Wind turbine acoustic noise. White Paper, Re-
newable Energy Research Laboratory, Department of Mechanical & Industrial Engineering,
University Massachusetts at Amherst, MA, 1003(January), 2002.
[54] Anthony L Rogers, James F Manwell, and Sally Wright. Wind turbine acoustic noise.
Renewable Energy Research Laboratory, Amherst: University of Massachusetts, 2006.
[55] Korn Saranyasoontorn and Lance Manuel. On the propagation of uncertainty in inow tur-
bulence to wind turbine loads. Journal of Wind Engineering and Industrial Aerodynamics,
96(5):503523, 2008.
[56] Korn Saranyasoontorn, Lance Manuel, and Paul S Veers. A comparison of standard coher-
ence models for inow turbulence with estimates from eld measurements. Transactions
of the ASME-N-Journal of Solar Energy Engineering, 126(4):10691082, 2004.
[57] E. Schmidt. Zur Theorie der linearen und nichtlinearen Integralgleichungen I. Teil: En-
twicklung willk urlicher Funktionen nach Systemen vorgeschriebener. Mathematische An-
nalen, 63(4):433476, 1907.
[58] D W Scott. Multivariate density estimation: theory, practice, and visualization. John
Wiley & Sons Inc, 1992.
[59] Mahmood M Shokrieh and Roham Raee. Simulation of fatigue failure in a full composite
wind turbine blade. Composite Structures, 74(3):332342, 2006.
101
[60] B W Silverman. Density estimation for statistics and data analysis. Chapman and Hall,
London, 1986.
[61] Eric Simley, Lucy Y Pao, Neil Kelley, Bonnie Jonkman, and Rod Frehlich. Lidar wind
speed measurements of evolving wind elds. In Proc. AIAA Aerospace Sciences Meeting,
2012.
[62] Georey Ingram Taylor. The spectrum of turbulence. Proceedings of the Royal Society of
London. Series A-Mathematical and Physical Sciences, 164(919):476490, 1938.
[63] S. W. Tsai and E. M. Wu. A general theory of strength for anisotropic materials. Journal
of Composite Materials, 5:5880, 1971.
[64] B A Turlach. Bandwidth Selection in Kernel Density Estimation: A Review. Technical
report, CORE and Institut de Statistique, 1993.
[65] U.S. Department of Defense. Military Handbook-MIL-HDBK-17-2F: Composite Materials
Handbook, Volume 2-Polymer Matrix Composites Materials Properties. june 2002.
[66] US DoE. 20% Wind Energy by 2030: Increasing wind energys contribution to US elec-
tricity supply. U.S. Department of Energy, Washington, DC, 2008.
[67] U.S. Energy Information Administration. Electric Power Monthly with Data for July 2012.
Technical report, U.S. Department of Energy, Washington, DC, 2012.
[68] Paul S Veers. Three-dimensional wind simulation. Technical report, Sandia National Labs.,
Albuquerque, NM (USA), 1988.
[69] D. Venturi. A fully symmetric nonlinear biorthogonal decomposition theory for random
elds. Physica D: Nonlinear Phenomena, 240(4-5):415425, February 2011.
[70] D. Venturi, X. Wan, and G. E. Karniadakis. Stochastic low-dimensional modelling of a
random laminar wake past a circular cylinder. Journal of Fluid Mechanics, 606:339367,
June 2008.
102
[71] David Verelst. Flexible wind turbine blades: a FEM-BEM coupled model approach. Tech-
nical report, Delft university of Technology, 2009.
[72] Theodore Von Karman. Progress in the statistical theory of turbulence. Proceedings of
the National Academy of Sciences of the United States of America, 34(11):530, 1948.
[73] Rolf W ustenhagen, Maarten Wolsink, and Mary Jean B urer. Social acceptance of renew-
able energy innovation: An introduction to the concept. Energy policy, 35(5):26832691,
2007.
[74] J. C. Wyngaard. Turbulence in the Atmosphere. Cambridge University Press, 2010.
[75] Y. Xie, J. Zola, and B. Ganapathysubramanian. FT-AdaGiO. 2013.
[76] D. Xiu and J.S. Hesthaven. High order collocation methods for the dierential equation
with random inputs. SIAM Journal on Scientic Computing, 27:11181139, 2005.
[77] D. Xiu and G. E. Karniadakis. The WienerAskey Polynomial Chaos for Stochastic Dif-
ferential Equations. SIAM Journal on Scientic Computing, 24(2):619, 2002.
[78] D. Xiu and G. E. Karniadakis. Modeling uncertainty in ow simulations via generalized
polynomial chaos. Journal of Computational Physics, 187(1):137167, May 2003.
[79] J. Zhang, S. Chowdhury, A. Messac, and L. Castillo. Multivariate and multimodal wind
distribution model based on kernel density estimation. ASME 2011 5th International Con-
ference on Energy Sustainability and 9th Fuel Cell Science, Engineering and Technology
Conference, August 2011.

S-ar putea să vă placă și